19
Astronomy & Astrophysics manuscript no. AA_2019_36093_p c ESO 2019 August 27, 2019 Spectral and orbital survey of medium-sized meteoroids ? Pavol Matloviˇ c 1 , Juraj Tóth 1 , Regina Rudawska 2 , Leonard Kornoš 1 , and Adriana Pisarˇ cíková 1 1 Faculty of Mathematics, Physics and Informatics, Comenius University, Bratislava, Slovakia e-mail: [email protected] 2 ESTEC/ESA, Keplerlaan 1, 2201 AZ Noordwijk, The Netherlands Received 2019 ABSTRACT Aims. We investigate the spectra, material properties, and orbital distribution of millimeter- to decimeter-sized meteoroids. Our study aims to distinguish the characteristics of populations of dierently sized meteoroids and reveal the heterogeneity of identified mete- oroid streams. We verify the surprisingly large ratio of pure iron meteoroids on asteroidal orbits detected among mm-sized bodies. Methods. Emission spectra and multi-station meteor trajectories were collected within the AMOS network observations. The sample is based on 202 meteors of -1 to -14 magnitude, corresponding to meteoroids of mm to dm sizes. Meteoroid composition is studied by spectral classification based on relative intensity ratios of Na, Mg, and Fe and corresponding monochromatic light curves. Helio- centric orbits, trajectory parameters, and material strengths inferred from empirical K B and P E parameters were determined for 146 meteoroids. Results. An overall increase of Na content compared to the population of mm-sized meteoroids was detected, reflecting weaker eects of space weathering processes on larger meteoroids. The preservation of volatiles in larger meteoroids is directly observed. We report a very low ratio of pure iron meteoroids and the discovery of a new spectral group of Fe-rich meteors. The majority of meteoroids on asteroidal orbits were found to be chondritic. Thermal processes causing Na depletion and physical processes resulting in Na-rich spectra are described and linked to characteristically increased material strengths. Numerous major and minor shower meteors were identified in our sample, revealing various degrees of heterogeneity within Halley-type, ecliptical, and sungrazing meteoroid streams. Our results imply a scattered composition of the fragments of comet 2P/Encke and 109P/Swift-Tuttle. The largest disparities were detected within α-Capricornids of the inactive comet 169P/NEAT and δ-Aquarids of the sungrazing 96P/Machholz. We also find a spectral similarity between κ-Cygnids and Taurids, which could imply a similar composition of the parent objects of the two streams. Key words. Meteorites, meteors, meteoroids; Minor planets, asteroids: general; Comets: general; techniques: spectroscopic 1. Introduction One of the ultimate goals of meteor spectroscopy is to reveal meteoroid composition from ground-based atmospheric obser- vations. This task is however quite dicult; the composition of the radiating plasma of a meteor does not fully reflect the chem- ical composition of the original meteoroid. This is particularly apparent for the refractory elements of Al, Ca, and Ti, which are lacking in the radiating plasma. Similarly, Si is a major compo- nent of most meteorites but is underrepresented in meteor spec- tra. There have been successful models of meteor radiation pro- viding relevant results of relative abundances of the main el- ements. These models have been mostly applied to individual bright fireball spectra captured in high resolution (Boroviˇ cka 1993; Trigo-Rodriguez et al. 2003; Ferus et al. 2018; Drouard et al. 2018). In order to survey larger samples of meteor spectra, a simplified method of spectral classification (Boroviˇ cka et al. 2005) has been introduced, which is also applicable to more ef- fective low-resolution video observations. The spectral classification method has since been applied by several authors. Within wider surveys, these studies were mainly focused on fainter meteors (Boroviˇ cka et al. 2005; Vojᡠcek et al. 2015, 2019). Other spectral analyses focused on individual me- teor showers (Boroviˇ cka 2010; Rudawska et al. 2014; Madiedo ? Tables 5 and 6 are only available in electronic form at the CDS via anonymous ftp to cdsarc.u-strasbg.fr (130.79.128.5) or via http://cdsweb.u-strasbg.fr/cgi-bin/qcat?J/A+A/ et al. 2014a; Matloviˇ c et al. 2017) or on linking specific spectral features with ablation properties (Bloxam & Campbell-Brown 2017; Vojᡠcek et al. 2019). The present work represents the widest spectral study of mid- sized meteoroids and also includes a large sample of orbital and structural properties, enabling a more complex view of the ana- lyzed meteoroid population. The presented results complement previous works focused on fainter meteors corresponding to ap- proximately mm-sized bodies (Boroviˇ cka 2001; Boroviˇ cka et al. 2005; Rudawska et al. 2014; Jenniskens et al. 2014), and bright fireballs often caused by meteoroids of over one meter in diame- ter that potentially break up and give rise to meteorites (Ceplecha et al. 1993; Trigo-Rodriguez et al. 2003; Madiedo et al. 2013). We aim to define the spectral properties of mm–dm sized mete- oroids and reveal the dierences compared to mm-sized bodies studied by Boroviˇ cka et al. (2005) and Vojᡠcek et al. (2015). From the fragments of asteroid 2008 TC 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009; Kohout et al. 2011). Here we try to find out how the material heterogeneities are reflected in low- resolution meteor spectra and to what degree they are exhibited within individual meteoroids from one parent object. Further- more, we focus on defining the characteristic spectral properties of meteoroids from dierent orbital sources (main belt, Jupiter- family, and Halley-type). It was suggested that while detailed composition (e.g., spe- cific types of chondrites) cannot be easily distinguished from Article number, page 1 of 19 arXiv:1908.01565v2 [astro-ph.EP] 26 Aug 2019

Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

Astronomy & Astrophysics manuscript no. AA_2019_36093_p c©ESO 2019August 27, 2019

Spectral and orbital survey of medium-sized meteoroids?

Pavol Matlovic1, Juraj Tóth1, Regina Rudawska2, Leonard Kornoš1, and Adriana Pisarcíková1

1 Faculty of Mathematics, Physics and Informatics, Comenius University, Bratislava, Slovakiae-mail: [email protected]

2 ESTEC/ESA, Keplerlaan 1, 2201 AZ Noordwijk, The Netherlands

Received 2019

ABSTRACT

Aims. We investigate the spectra, material properties, and orbital distribution of millimeter- to decimeter-sized meteoroids. Our studyaims to distinguish the characteristics of populations of differently sized meteoroids and reveal the heterogeneity of identified mete-oroid streams. We verify the surprisingly large ratio of pure iron meteoroids on asteroidal orbits detected among mm-sized bodies.Methods. Emission spectra and multi-station meteor trajectories were collected within the AMOS network observations. The sampleis based on 202 meteors of -1 to -14 magnitude, corresponding to meteoroids of mm to dm sizes. Meteoroid composition is studiedby spectral classification based on relative intensity ratios of Na, Mg, and Fe and corresponding monochromatic light curves. Helio-centric orbits, trajectory parameters, and material strengths inferred from empirical KB and PE parameters were determined for 146meteoroids.Results. An overall increase of Na content compared to the population of mm-sized meteoroids was detected, reflecting weaker effectsof space weathering processes on larger meteoroids. The preservation of volatiles in larger meteoroids is directly observed. We reporta very low ratio of pure iron meteoroids and the discovery of a new spectral group of Fe-rich meteors. The majority of meteoroidson asteroidal orbits were found to be chondritic. Thermal processes causing Na depletion and physical processes resulting in Na-richspectra are described and linked to characteristically increased material strengths. Numerous major and minor shower meteors wereidentified in our sample, revealing various degrees of heterogeneity within Halley-type, ecliptical, and sungrazing meteoroid streams.Our results imply a scattered composition of the fragments of comet 2P/Encke and 109P/Swift-Tuttle. The largest disparities weredetected within α-Capricornids of the inactive comet 169P/NEAT and δ-Aquarids of the sungrazing 96P/Machholz. We also find aspectral similarity between κ-Cygnids and Taurids, which could imply a similar composition of the parent objects of the two streams.

Key words. Meteorites, meteors, meteoroids; Minor planets, asteroids: general; Comets: general; techniques: spectroscopic

1. Introduction

One of the ultimate goals of meteor spectroscopy is to revealmeteoroid composition from ground-based atmospheric obser-vations. This task is however quite difficult; the composition ofthe radiating plasma of a meteor does not fully reflect the chem-ical composition of the original meteoroid. This is particularlyapparent for the refractory elements of Al, Ca, and Ti, which arelacking in the radiating plasma. Similarly, Si is a major compo-nent of most meteorites but is underrepresented in meteor spec-tra. There have been successful models of meteor radiation pro-viding relevant results of relative abundances of the main el-ements. These models have been mostly applied to individualbright fireball spectra captured in high resolution (Borovicka1993; Trigo-Rodriguez et al. 2003; Ferus et al. 2018; Drouardet al. 2018). In order to survey larger samples of meteor spectra,a simplified method of spectral classification (Borovicka et al.2005) has been introduced, which is also applicable to more ef-fective low-resolution video observations.

The spectral classification method has since been applied byseveral authors. Within wider surveys, these studies were mainlyfocused on fainter meteors (Borovicka et al. 2005; Vojácek et al.2015, 2019). Other spectral analyses focused on individual me-teor showers (Borovicka 2010; Rudawska et al. 2014; Madiedo

? Tables 5 and 6 are only available in electronic form at theCDS via anonymous ftp to cdsarc.u-strasbg.fr (130.79.128.5) or viahttp://cdsweb.u-strasbg.fr/cgi-bin/qcat?J/A+A/

et al. 2014a; Matlovic et al. 2017) or on linking specific spectralfeatures with ablation properties (Bloxam & Campbell-Brown2017; Vojácek et al. 2019).

The present work represents the widest spectral study of mid-sized meteoroids and also includes a large sample of orbital andstructural properties, enabling a more complex view of the ana-lyzed meteoroid population. The presented results complementprevious works focused on fainter meteors corresponding to ap-proximately mm-sized bodies (Borovicka 2001; Borovicka et al.2005; Rudawska et al. 2014; Jenniskens et al. 2014), and brightfireballs often caused by meteoroids of over one meter in diame-ter that potentially break up and give rise to meteorites (Ceplechaet al. 1993; Trigo-Rodriguez et al. 2003; Madiedo et al. 2013).We aim to define the spectral properties of mm–dm sized mete-oroids and reveal the differences compared to mm-sized bodiesstudied by Borovicka et al. (2005) and Vojácek et al. (2015).

From the fragments of asteroid 2008 TC3 we know thatmeter-scale meteoroids can be compositionally heterogeneous(Jenniskens et al. 2009; Kohout et al. 2011). Here we try tofind out how the material heterogeneities are reflected in low-resolution meteor spectra and to what degree they are exhibitedwithin individual meteoroids from one parent object. Further-more, we focus on defining the characteristic spectral propertiesof meteoroids from different orbital sources (main belt, Jupiter-family, and Halley-type).

It was suggested that while detailed composition (e.g., spe-cific types of chondrites) cannot be easily distinguished from

Article number, page 1 of 19

arX

iv:1

908.

0156

5v2

[as

tro-

ph.E

P] 2

6 A

ug 2

019

Page 2: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

A&A proofs: manuscript no. AA_2019_36093_p

Fig. 1: Slovak part of the global AMOS network. Observationsof the AMOS-Spec and AMOS stations by the Astronomical andGeophysical Observatory in Modra (AGO), Arborétum TesárskeMlynany (ARBO), Kysucké Nové Mesto (KNM), and Važec(VAZ) were used in this work. Red, green, and blue labels des-ignate operating standard AMOS cameras, spectral stations, andplanned stations.

low-resolution meteor spectra (Borovicka et al. 2015), the roughcomposition (chondritic, achondritic, metallic) can be identified.We attempt to verify this hypothesis and try to look for additionalinformation about the meteoroid nature and structure that can berevealed from determined orbital and atmospheric parameters.

One of the most surprising results of the study by Borovickaet al. (2005) was the large number of the detected pure-ironmeteoroids, suggesting that iron meteoroids prevail among mm-sized meteoroids on asteroidal orbits. These results are not con-sistent with the expectations based on what we know about thetransport of large main-belt asteroids (MBA) to near-Earth or-bit (NEO) space (Granvik et al. 2017). Our work provides moredetail into the spectral properties of mm- to dm-sized asteroidalmeteoroids, which could help to clarify this inconsistency.

2. Methods and observations

The meteor data analyzed in this work come from observationsof the All-sky Meteor Orbit System (AMOS) network devel-oped and operated by the Comenius University in Bratislava.The network consists globally of 11 standard AMOS systemsand 5 spectral AMOS-Spec systems. The stations are positionedat different locations in Slovakia, the Canary Islands, Chile, andHawaii. Standard AMOS systems provide continuous all-sky de-tection of meteors and are used to determine the atmospheric tra-jectory and original heliocentric orbit of observed meteoroids.In this work, we use the observations of the Slovak part of theAMOS network (Fig. 1). All of the spectra presented here wereobserved by the AMOS-Spec station at the Modra Observatory.More information about the AMOS network can be found inTóth et al. (2015, 2019) and Zigo et al. (2013).

The presented spectra span the entire year and include spo-radic meteors as well as meteors from known meteoroid streams.All spectra with S/N > 4 per pixel were selected for our analy-sis. In some cases, poor atmospheric conditions (clouds, strongMoon illumination) and acute entry angle hampered the qual-ity of collected spectra. Approximately 10% of collected spectrawere excluded from this analysis due to poor quality. Further-more, in cases of bright meteor flares, only frames without sig-nificant saturation were used for processing.

Fig. 2: Example of the zero-order image of meteorM20171220_223458 (trail on the right) with the first-orderspectrum (distributed trails on the left) captured by theAMOS-Spec station in AGO Modra.

2.1. AMOS instrumentation

The standard AMOS system consists of four major components:a fish-eye lens, an image intensifier, a projection lens, and a digi-tal video camera. The resulting field of view (FOV) of AMOS is180 x 140 with an image resolution of 8.4 arcmin/pixel, anda video frame rate of 15 fps. Limiting magnitude for stars isabout +5 mag for a single frame, and the detection efficiencyis lower for moving objects, approximately +4 mag at typicalmeteor speeds due to the trailing loss.

The AMOS-Spec is a semi-automatic remotely controlledvideo system for meteor spectra observations. The display com-ponents of the AMOS-Spec consist of a 30 mm f/3.5 fish-eyelens, an image intensifier (Mullard XX1332), a projection lens(Opticon 1.4/19 mm), and a digital camera (Imaging SourceDMK 51AU02). This setup yields a 100 circular FOV (Fig.2) with a resolution of 1600 x 1200 pixels and a frame rate of12/s. To our knowledge, AMOS-Spec is one of the largest FOVspectrographs used for meteor observations. Diffraction of theincoming light is provided by a blazed holographic grating with1000 grooves/mm placed above the lens. The resulting disper-sion varies due to the geometry of the all-sky lens with an aver-age value of 1.3 nm/px. This translates to a spectral resolution ofapproximately R = 200. The limiting magnitude of the AMOS-Spec for meteors (zero order) is similarly to that of the standardAMOS, namely around +4 mag, while the limiting magnitudefor a spectral event (first order) is -1 mag. The system covers thewhole visual spectrum range of 370–900 nm with a sensitivitypeak at 470 nm. An example of an spectral event captured by theAMOS-Spec can be seen in Fig. 2.

2.2. Spectral reduction

The image processing and reduction of spectra analyzed in thiswork follow the procedure described in Matlovic et al. (2017).In the first step, all spectral recordings are corrected for darkframe, flat-field images and have the star background image sub-

Article number, page 2 of 19

Page 3: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

Pavol Matlovic et al.: Spectral and orbital survey of medium-sized meteoroids

0.0

0.2

0.4

0.6

0.8

1.0

350 400 450 500 550 600 650 700 750 800 850 900Wavelength [nm]

Rel

ativ

e re

spon

se

Fig. 3: Spectral response curve of the AMOS-Spec system.

tracted to reduce noise and other sources of illumination. Theall-sky geometry of the AMOS-Spec lenses causes slight curva-ture of events captured near the edge of the FOV. Each spectrumis therefore manually scanned in the ImageJ1 program on indi-vidual frames with spectrum signal. The width of the scanningslit is adapted to the spectrum geometry.

All spectra are first scaled based on recognized spectral linesand polynomial fit of second or higher order. The wavelengthscale is later fine-tuned during the fitting procedure. The spec-trum is corrected for the spectral sensitivity of the system (Fig.3). The spectral response curve was determined using a calibra-tion halogen lamp measured in a laboratory environment. The at-mospheric transmittance at different stations also affects the spe-cific spectral sensitivity, mainly in the near-UV region. Transmit-tance differences of < 2% (Bertaux et al. 2014) can be neglectedin the 510-600 nm range of wavelengths under consideration.

Each spectrum was fitted with the most significant spectrallines (low temperature, high temperature, atmospheric, and wakelines) in meteors using the Fityk software (Wojdyr 2010)2. Themodeled lines were used to create a synthetic spectrum whichwas then fit to the measured and calibrated meteor spectrum. Thefitting procedure follows the Levenberg–Marquardt algorithm,also known as the damped least-squares method. The modeledlines have Gaussian profiles with a full width at half maximum(FWHM) of typically 3 nm. Moreover, the most notable molec-ular N2 bands of the first positive system present in the meteorspectra were fitted using the relative intensities from Borovickaet al. (2005), adjusted for our spectra. The atmospheric emissionlines of O I, N I, and N2 bands were fitted in the synthetic spec-trum depending on the meteor speed and were subtracted beforethe measurement of meteoroid emission lines.

In this work, we focus on the main meteor multiplets of MgI – 2, Na I – 1, and Fe I – 15, which form the basis of the spectralclassification (Borovicka et al. 2005). The intensities of thesemultiplets were measured in the fitted synthetic spectrum aftersubtraction of the continuous radiation and atmospheric emis-sion. The resulting ratios were then applied for the spectral clas-sification. The modeled contributions of all recognized lines ofthe Fe I - 15 multiplet were summed. These multiplets were se-lected as they can be accurately measured in most meteor spec-tra, and are all in the region of high instrumental spectral sen-

1 https://imagej.nih.gov/ij/2 http://fityk.nieto.pl/

Fig. 4: Fit of the synthetic spectrum (yellow) on measured cal-ibrated meteor spectrum (green) as a sum of the main emissioncontributions (red) in the 510-600 nm region. Residuals of the fitcan be seen below the spectrum.

sitivity. Nevertheless, the measurement of these lines must betaken with caution, because the Na multiplet overlaps with a se-quence of N2 bands, while the wake lines of Fe can influencethe intensity of the Mg line. All these effects were accounted forin the reduction process. The errors of determined Na/Mg andFe/Mg ratios presented here were determined from the signal-to-noise ratio (S/N) analysis for individual lines.

It must be noted that the spectral classification procedure wasoriginally developed for fainter meteors in the magnitude range+3 to -1, corresponding to meteoroid sizes of 1 - 10 mm. Oursystem observes meteor spectra of -1 up to -14 magnitude, cor-responding to meteoroid sizes of a few mm up to a few decime-ters. We expect that the same physical conditions fitted by thethermal equilibrium model applied for bright photographic me-teors (Borovicka 1993) as well as for fainter meteors (Borovickaet al. 2005) can also be assumed for the meteoroid populationobserved by our system. The effect of self-absorption in brightermeteor spectra was examined in individual cases. Individualframes with saturated meteor spectra were excluded from theanalysis.

Figure 4 shows an example of a meteor fit in the 510 - 600nm region. In specific cases, the positions of atypical lines weresearched for in the NIST Atomic Spectra Database (Kramidaet al. 2018) and in the Astrophysical Multiplet Table (Moore1945). Given the large scope of this work, we do not addressin detail the minor and atypical emission lines detected duringthe fitting procedure. The data set analyzed here however can beused for further investigations.

2.3. Astrometry and photometry

Out of the 202 meteors analyzed here, 146 were observed bymultiple stations of the AMOS network. Triangulation of meteorpaths observed from different stations enable us to accuratelydetermine the atmospheric trajectory and original heliocentricorbit of the incoming meteoroid. Meteors captured by multiplestations, along with their spectra, form the ground base for thiswork.

The UFOCapture detection software (SonotaCo 2009) wasused during real-time observations to detect and record all of thestudied meteors. Star and meteor coordinates from each frame ofeach event were measured using the original AMOS software. Infuture, the AMOS software is also intended to replace the UFO-

Article number, page 3 of 19

Page 4: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

A&A proofs: manuscript no. AA_2019_36093_p

Capture software for the real-time detection and positional mea-surements at all AMOS stations.

Star identification and astrometry was performed using theRedSky software (upgraded from an earlier version of Kornošet al. 2018) based on the all-sky procedure of Borovicka et al.(1995). RedSky uses mathematical transformation for comput-ing rectangular coordinates of catalog stars that are comparedwith measured stars. The identified stars are used to determinethe 13 plate constants from which the zenith distance and az-imuth is computed. As a reference, the Sky Catalogue 2000.0 isused up to +6th magnitude including the color index. The pre-cision of the AMOS astrometry is on the order of 0.02 - 0.03,which translates to an accuracy of tens of meters for atmosphericmeteor trajectory.

Meteor orbits were determined using the Meteor Trajectory(MT) software based on the procedure of Ceplecha (1987). Theearly version of the program has shown promising results interms of precision (Kornoš et al. 2015, 2018) and has recentlybeen upgraded for new functionalities. The outputs of the or-bital measurements have been, for individual cases, tested andvalidated on shared observations with the Astronomical Instituteof the Czech Academy of Sciences (Dr. Pavel Spurný, privatecommunication). The program computes orbital and geophysi-cal parameters together with their uncertainties based on a MonteCarlo simulation within an uncertainty in the radiant position andgeocentric velocity.

Meteor photometry is also performed within the MT pro-gram, providing apparent and absolute magnitudes and corre-sponding light curves. The precision of initially obtained mag-nitudes was further examined individually by visual calibrationbased on comparison with bright stars, planets, and for brightmeteors with the Moon in different phases. Absolute magnitudeswere then determined by correction to a standard altitude of 100km at the observation zenith, and for atmospheric extinction.

The photometry of the AMOS system is not well adaptedto exceptionally bright fireballs (< -8 mag). Based on previousexperiences, we suspect that the magnitudes of the brightest fire-balls in our sample might be underestimated (Dr. Pavel Spurný,private communication). The underestimation mainly concernsfireballs with bright flares, in which the sharp peaks of brightnessare not detected from the integrated video photometry. There areseveral such cases in our sample. For these, the uncertainty onthe determined magnitude and the consequent photometric pa-rameters is higher. The majority of meteors presented here ex-hibited brightnesses typically between -2 mag and -7 mag (Fig.5). In these cases, saturation did not significantly affect the pho-tometric calibration. Monochromatic light curves were resolvedby our own code based on the relative intensities of the mainmeteor emission lines (Mg I, Na I, Fe I) at different atmosphericheights.

3. Spectral classification

Overall 202 meteor spectra captured by the AMOS-Spec sys-tem during the time period between December 2013 and August2017 were used for the spectral analysis. The sample includesmeteors calculated to be from -1 to -11 mag (Fig. 5). As notedin section 2.3, some of the magnitudes of the brightest fireballsin our sample (< -8 mag) might be underestimated. We estimatethat the lower end of the studied magnitude range might reachup to -14 mag. The magnitude range corresponds to meteoroidsin the mm to dm size range.

The spectral classification of our sample based on the relativeintensities of Mg I - 2, Na I - 1, and Fe I - 15 is in Fig. 6. Spec-

0

10

20

30

−12 −11 −10 −9 −8 −7 −6 −5 −4 −3 −2 −1Absolute magnitude

Cou

nt

0

5

10

15

20

−12 −11 −10 −9 −8 −7 −6 −5 −4 −3 −2 −1

Cou

nt

Tisserand type

ASTHTJF

Absolute magnitude

Fig. 5: Histogram of absolute magnitudes of meteors in our sam-ple (upper) and magnitude distribution relative to the type of he-liocentric orbit based on the Tisserand parameter (lower). Thefollowing abbreviations are used: AST for asteroidal orbits, JFfor Jupiter-family type orbits, and HT for Halley-type cometaryorbits. Only magnitudes of meteors observed by multiple sta-tions are displayed.

tra were classified in accordance with the definitions introducedby Borovicka et al. (2005). Out of the 202 measured meteors,145 have been identified as normal type, typically positioned inthe middle part of the ternary diagram. Normal-type meteors aredefined as those lying close to the expected position for chon-dritic bodies, as modeled by Borovicka et al. (2005) and visibleas the chondritic curve in Fig. 6b). These theoretical values andthe resulting mean curve were determined assuming chondriticcomposition and range of temperatures, densities, and sizes ofthe radiating plasma. This curve might not accurately representthe characteristic chondritic area within our sample of meteorscreated typically by larger meteoroids. It is likely that the con-ditions of mean temperature and size of the radiating plasma areshifted in our population. While no theoretical simulation wasperformed, the characteristic positions of normal-type meteorswere simply inferred from the densest part of the distribution inFig. 6. Normal-type spectra form the largest group of our sample(Fig. 7), similarly to previous studies of mm-sized meteoroids(Borovicka et al. 2005; Vojácek et al. 2015, 2019).

Normal meteoroids are along with Na-poor, Na-enhanced,and Fe-poor meteoroids sometimes referred to as mainstreammeteoroids. Na-poor meteors are defined as having a weaker-than-expected (but still visible) Na line for their given speed. Na-enhanced meteoroids have an obviously stronger-than-expectedNa line for their given speed but not so dominant as in Na-richmeteoroids. Fe-poor have the expected Na/Mg ratio but Fe linesthat are too faint to be classified as normal. More distinct or so-called nonmainstream classes include three groups: Iron meteorsare defined as having spectra dominated by Fe lines. Na-free me-teors show no Na line but are not classified as Irons. Finally, Na-rich meteors have spectra dominated by the Na line, with Na/Mg

Article number, page 4 of 19

Page 5: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

Pavol Matlovic et al.: Spectral and orbital survey of medium-sized meteoroids

20

40

60

80

100

20

40

60

80

100

20 40 60 80

Fe

Mg

Na

F I ­ 15e

I ­ 2Mg I ­ 1Na

Fe poorFe richIronsNa enhancedNa freeNa poorNa richNormal

(a)

20

40

60

80

100

20

40

60

80

100

20 40 60 80

Fe

Mg

Na

F I ­ 15e

I ­ 2Mg I ­ 1Na

Fe poorIronsNa enhancedNa freeNa poorNa richNormal

40 30 20

15

(b)

Fig. 6: Ternary diagram displaying the spectral classification of 202 meteors of -1 to -14 mag (corresponding to meteoroids of mmto dm sizes) observed by the AMOS-Spec system during 2013-2017 (a) and (b) comparison to the spectral classification of +3 to -1mag meteors (corresponding to meteoroid sizes of 1-10 mm) based on data from Borovicka et al. (2005), including a curve showingthe expected range for chondritic composition as a function of meteor speed (in km s-1), as modeled by Borovicka et al. (2005).

and Na/Fe ratios obviously higher than expected for chondriticbodies (Borovicka et al. 2005).

Besides the determined Na/Mg/Fe intensity ratios, the effectof speed dependence (discussed in Section 4) was taken into ac-count for the classification of Na-enhanced and Na-poor mete-ors. Furthermore, the effects of saturation, self-absorption, andintensities of other Fe multiplets were taken into account forclassification of Fe-rich meteors. This is the cause of the partialoverlap of spectral classes in Fig. 6.

The shift of the spectral classification among two popula-tions of differently sized meteoroids can be seen by comparisonwith the ternary diagram in Fig. 6 of Borovicka et al. (2005).A similar comparison can be made with the ternary diagramsfrom Vojácek et al. (2015) or Vojácek et al. (2019), which alsodeal with mm-sized meteoroids and show equivalent distribu-tions. The increase of the Na/Mg ratio in spectra of larger mete-oroids is also apparent from Fig. 8.

We identified three possible explanations for this effect. Thefirst is instrumental: The spectral sensitivity curve of the AMOS-Spec system shows lower efficiency at 589.2 nm, the peak wave-length of the Na I doublet. We checked the accuracy of the sensi-tivity curve previously determined from a reference Jupiter spec-trum. A new, updated response curve was remeasured and con-firmed by several calibration lamps in the laboratory. The re-sponse curve error is not sufficient to explain such a shift.

The second possible explanation is physical. The differencesbetween the magnitude ranges of studied meteor populations arelikely to affect the intensity of the Na line. It has been noted thatthe Na line is often observed earlier (at higher altitudes) thanother lines. This relates to the low excitation potential of the NaI line, which can radiate at lower temperatures compared to MgI. This effect would however favor the increase of Na/Mg ratio in

fainter meteors, compared to our sample caused by meteoroidsof larger size. The opposite is observed and, as we discuss furtherbelow, this effect is much more dependent on meteor speed.

Finally, we suggest that the observed shift does in fact re-flect the composition of meteoroids. Sodium is a volatile ele-ment, which can be easily depleted from small bodies by spaceweathering processes. These effects mainly concern the exposureto solar radiation (and specifically thermal desorption when me-teoroids are in close proximity to the Sun) and probably the solarwind and cosmic ray irradiation. Depending on the compositionand time of exposure, these processes affect mainly the outer lay-ers of interplanetary bodies. For smaller meteoroids, the spaceweathering processes likely cause more effective volatile deple-tion within the meteoroid volume. In larger meteoroids studiedin our sample, volatiles can be partially depleted from the outerlayers but can remain intact below the surface.

The effect of Na depletion might not only concern the spaceweathering processes before entry into Earth’s atmosphere. Thepreheating of the meteoroid in Earth’s atmosphere could causesome degree of Na depletion before the meteor spectrum canbe observed. The preheating would also more significantly af-fect smaller meteoroids. Some degree of Na enhancement in 13larger meteoroids was already noted by Trigo-Rodriguez et al.(2003). The effect is also confirmed by the apparent differencesbetween the number of Na-poor and Na-free meteors observedin our sample compared to that of Borovicka et al. (2005). Only11.5% of our sample (Fig. 7) represents these classes, comparedto 34% among mm-sized meteoroids.

The second most notable difference between the two me-teoroid populations is the apparent lack of iron bodies in oursample. Meteor spectra dominated by Fe lines are assumed tocome from meteoroids composed mainly of iron–nickel alloy

Article number, page 5 of 19

Page 6: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

A&A proofs: manuscript no. AA_2019_36093_p

25

50

75

Normal

Na poor

Na enhanced

Fe poor Fe rich

Na free Na rich

Iron

Fig. 7: Distribution of spectral classes identified within the sam-ple of 202 meteoroids in the mm to dm size range.

(Borovicka et al. 2005; Capek & Borovicka 2017). Only oneiron meteor has been identified in our sample (0.5%), comparedto approximately 14% among the fainter meteors in the sam-ple of Borovicka et al. (2005). Similarly, 10% of the sample de-tected by Vojácek et al. (2019) were iron meteoroids. This isan unexpectedly large proportion considering the statistics fromknown meteorite falls. According to the Meteoritical BulletinDatabase3, only 3.7% of all meteorites have been identified asiron meteorites. Meteorite samples only include strong asteroidalbodies, which are able to withstand the atmospheric flight. Theproportion of iron bodies in interplanetary space would be ex-pected to be even lower. According to the results of Borovickaet al. (2005), iron meteoroids are dominant among meteoroidson asteroidal orbits. This is in strong contrast to our results, sug-gesting that most cm to dm sized asteroidal meteoroids are chon-dritic. Our results are consistent with the expectations based onwhat we know about the transport of large main-belt asteroids toNEO space (Granvik et al. 2017).

Capek & Borovicka (2017) pointed out the specific abla-tion process of small (0.7 - 2.1 mm) iron meteoroids. The lightcurves typically exhibit very steep initial increase of brightnessand a gradual decrease. Similar light curves have been detectedamong small refractory meteoroids by Campbell-Brown (2015),likely also composed of iron–nickel alloy or iron sulfide grains.According to Capek & Borovicka (2017) and Girin (2017), theablation of these meteoroids is in the form of droplets releasedfrom a liquid layer at the meteoroid surface. Borovicka et al.(2005) speculated that if the meteoroid consists of metallic ironwith high thermal conductivity, the whole mm-sized meteoroidcould be melted completely before the vaporization. The largeproportion of iron meteoroids in the mm-sized population couldbe related to specific formation processes of these bodies andtypical sizes of iron–nickel grains. Further investigation is nec-essary to reveal the processes by which the population of smalliron meteoroids is formed.

While the abundance of iron bodies in our sample is verylow, we detected several meteors positioned between the normaltype and iron class. We introduce a new spectral class of Fe-rich meteors. These are defined as having significantly higherFe/Mg intensity ratio compared to normal-type meteors, while

3 https://www.lpi.usra.edu/meteor/

−1

0

1

2

10 20 30 40 50 60 70Meteor speed [km/s]

Na/

Mg

(log 1

0)

Fe poorFe richIronsNa enhancedNa freeNa poorNa richNormal

Fig. 8: Observed Na/Mg intensity ratio as a function of meteorspeed. The solid line represents a cubic fit to the running averagewith standard error of the mean for meteors classified as normal.The dashed line represents the fit of normal-type meteors fromBorovicka et al. (2005) dealing with a mm-sized meteoroid pop-ulation.

still showing a notable presence of both Na I and Mg I. Thesemeteors are positioned in the upper part of the ternary diagram(Fig. 6). No meteors were detected in this part of the ternary di-agram in the surveys of fainter meteors (Borovicka et al. 2005;Vojácek et al. 2015, 2019). In some cases, the intensity of Fe canbe overestimated in meteors during bright flares. The pixel sat-uration and optical thickness of the radiating plasma can causean apparent increase of the Fe/Mg ratio. To reduce this effect,frames with notably saturated spectra were neglected from theline intensity measurements. Furthermore, the Fe/Mg ratio wasspecifically studied in nonsaturated frames before and after me-teor flares, and the relative intensity of the Fe I - 41 multiplet inthe 438 - 441 nm region was evaluated to identify Fe-rich mete-ors.

4. Meteor orbits

By triangulation of meteors observed by multiple stations, wedetermined atmospheric trajectories and heliocentric orbits for146 of the 202 meteors analyzed in this work. The correspondingparameters for all multi-station meteors can be found in Tables5 and 6 (available at the CDS). In most cases, the meteor entryspeed was determined with an uncertainty of < 1 km s-1. Hy-perbolic solutions were obtained for 15 meteors. In such cases,some orbital elements (a, e, TJ and Q) were neglected from Ta-ble 5 (available at the CDS).

The dependence of the Na/Mg ratio on meteor speed isshown in Fig. 8. This also illustrates the previously noted higherNa/Mg values for larger meteoroids measured by us comparedto the ratio in smaller meteoroids studied by Borovicka et al.(2005). The Na/Mg intensity ratio is dependent on meteor speed(temperature) as a result of the low excitation of Na I line (2.1eV) compared to Mg I (5.1 eV). Among mm-sized particles(Borovicka et al. 2005), this effect is observed for meteor speedsbelow 40 km s-1. Similar functionality is detected in our sampleof larger meteoroids, with the threshold value of meteor speedbetween 35 and 40 km s-1 (Fig. 8). The breaking point near 35km s-1 was recently found as a good fit for data of Vojácek et al.(2019).

Article number, page 6 of 19

Page 7: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

Pavol Matlovic et al.: Spectral and orbital survey of medium-sized meteoroids

0

50

100

150

−1 0 1 2 3 4 5 6Tisserand parameter

Incl

inat

ion

[°]

Fe poorFe richIronsNa enhancedNa freeNa poorNa richNormal

Fig. 9: Inclination vs. Tisserand parameter plot for measured me-teors of different spectral classes. The colored areas representasteroidal (green) and Jupiter-family-type (red) orbits.

The orbits of meteors were classified based on the Tisserandparameter relative to Jupiter. The parameter is often used to de-fine the orbital origin of interplanetary bodies (Kresak 1979),particularly comets, and can be in the same way used for mete-oroids. We use it here to distinguish between asteroidal (AST)orbits (TJ > 3), typical for bodies originating in the main aster-oid belt, Jupiter-family type (JT) orbits (2 < TJ < 3 and i < 45)of bodies from the Kuiper Belt captured by the gravitational in-fluence of Jupiter, and Halley-type (HT) orbits (TJ < 2 and i >45) typical for cometary bodies originating in the Oort cloud.Figure 5 displays a distribution of meteor magnitudes for differ-ent types of meteoroid orbits based on the Tisserand parameter.

Overall 61 meteoroids were found on Halley-type orbits, 42meteoroids on Jupiter-family type orbits, and 28 meteoroids onasteroidal orbits (Fig. 9). To some degree, the orbital origin ofmeteoroids is reflected in the detected spectral properties of me-teors. Figure 10 shows the increase of Fe/Mg intensity ratio formeteoroids originating in the main asteroid belt and decrease ofFe/Mg intensity ratio for meteoroids on Halley-type orbits. Weassume that the majority of meteoroids on orbits with TJ > 3(with the main exception of several meteoroids from the Tauridstream) are of asteroidal origin.

These chondritic bodies have a larger iron content in generalcompared to mostly cometary meteoroids on Halley-type orbits.The distinction between Jupiter-family and Halley-type mete-oroids in the characteristic Fe/Mg ratios is less apparent, thoughon average is also present. This is mainly caused by the presenceof Na-free meteoroids in this region and by numerous bodiesnear the TJ = 3 limit. The mean Fe/Mg intensity ratio (± stan-dard error of the mean) for meteoroids on asteroidal orbits (TJ >

3) is Fe/MgAS T = 2.81 ± 0.33, for Jupiter-family meteoroidsis Fe/MgJF = 1.28 ± 0.12, and for Halley-type meteoroids isFe/MgHT = 1.01 ± 0.06. The larger standard error of Fe/Mg forasteroidal orbits is mainly caused by the presence of short-periodcometary meteoroids, mainly Taurids and α-Capricornids.

The Na/Mg ratio apparently increases for meteoroids on as-teroidal orbits (Fig. 10) with Na/MgAS T = 4.29 ± 1.21 comparedto Na/MgJF = 2.31 ± 0.34 and Na/MgHT = 1.68 ± 0.10. Thelarger standard errors emphasize the wide spread of materials on

−1.0

−0.5

0.0

0.5

1.0

1.5

−1 0 1 2 3 4 5 6Tisserand parameter

Fe/

Mg

(log 1

0)

Fe poorFe richIronsNa enhancedNa freeNa poorNa richNormal

−1

0

1

2

−1 0 1 2 3 4 5 6Tisserand parameter

Na/

Mg

(log 1

0)

Fe poorFe richNa enhancedNa freeNa poorNa richNormal

Fig. 10: Observed Fe/Mg intensity ratio (upper) and Na/Mg in-tensity ratio (lower) as a function of the Tisserand parameter rel-ative to Jupiter. A cubic fit to the running average of all datapoints and the standard error of the mean are plotted. The coloredareas represent asteroidal (green), Jupiter-family type (pink), andHalley-type (white) orbits.

Jupiter-family and asteroidal orbits. Still, the observed effect ismainly caused by an increase of the Na/Mg ratio in slower me-teors (Fig. 8) characteristic of asteroidal meteoroids. All of themeteors with TJ > 3 have initial velocities vi < 35 km s-1.

Figures 9 and 11 both show that the distinction between as-teroidal and cometary bodies near the TJ = 3 boundary is dif-ficult. There is indeed a large overlap between asteroids andJupiter-family comets near TJ = 3. Therefore, additional struc-tural and spectral data must be considered to distinguish betweenthe two sources. Meteoroid streams positioned near the TJ = 3boundary are sometimes defined as ecliptical meteor showers.From the showers identified in our sample, we find the Tauridcomplex showers, the α-Capricornids, the η-Virginids, and theκ-Cygnids. Meteoroids with TJ >> 3 are likely chondritic. Thissuggests that all Na-rich meteoroids in our sample are of aster-oidal material.

Orbits with q < 0.2 au have been classified separately bysome authors as Sun-approaching. This group distinguishes me-teoroids with small perihelion distances, in which intense solarradiation affects the spectra. The distinction is displayed on aperihelion versus aphelion plot (Fig. 11). A cluster of meteoroidscan be observed near Q = 4.5 au, formed by meteors from eclip-

Article number, page 7 of 19

Page 8: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

A&A proofs: manuscript no. AA_2019_36093_p

1

10

100

1000

0.01 0.10 1.00 10.00Perihelion [au]

Aph

elio

n [a

u]

Fe poorFe richIronsNa enhancedNa freeNa poorNa richNormal

Fig. 11: Perihelion vs. aphelion plot in logarithmic scale forall observed meteoroids. The colored areas represent asteroidal(green), cometary (pink), and Sun-approaching orbits (yellow).

tical showers (mainly the Taurids) affected by the the gravita-tional influence of Jupiter. The aphelion of the orbit of Jupiter issometimes used as a boundary between comets and asteroids.

5. Shower meteors

Of the 146 meteoroids for which orbital data is available, 80were identified as members of known meteor showers and 66were sporadic. In general, sporadic meteors are thought to havebeen released from their parent body thousands to millions ofyears ago and to have moved separately from the recognizedparent stream. The shower association was performed withinthe MT software based on radiant position, meteor speed, andsolar longitude. The orbital similarity was evaluated using theSouthworth-Hawkins criterion (Southworth & Hawkins 1963)which defines the difference between two meteors representedas a point in a five-dimensional phase space of orbital elements.

We considered both established and working list streamsof the IAU Meteor Data Center (MDC) (Jopek & Kanuchová2017) for the identification. In cases where multiple streamswere found close to the meteor orbit, we individually estimatedthe most probable source considering the orbital similarity, radi-ant position, and shower activity.

Established showers are caused by confirmed meteoroidstreams with well-defined meteor activity. The spectral classi-fication of meteors associated with established Jupiter-family-type or ecliptical and Halley-type meteoroid streams is shownin Fig. 12. Overall, the Jupiter-family and ecliptical showers areshifted slightly to the right in the ternary diagram compared toHalley-type showers. This mainly reflects the generally lowermeteor speeds of their meteors. The ternary diagrams in Fig. 12are dominated by Perseids and Taurids, respectively. All of theTaurid meteoroids analyzed here were observed during the 2015outburst and were analyzed previously by Matlovic et al. (2017).

5.1. Halley-type

The most represented meteoroid stream in our sample are thePerseids originating in comet 109P/Swift–Tuttle. A list of me-teors associated with this stream and including their orbits andatmospheric data is provided in Table 1. Several more Perseidspectra were observed, but with the only recognizable lines being

atmospheric O I and N I. These types of spectra were excludedfrom our analysis. Most Perseid meteors are positioned in themiddle of the ternary diagram (Fig. 12) with normal-type spec-tra close to the expected chondritic ratios of Na/Mg/Fe. A rela-tively large variation of Fe content is observed, between 10 and40% in the ternary ratios. Four Fe-poor Perseids are observed.The Na/Mg ratios for Perseids are within the expected normal-type region, with one exception identified as Na-enhanced. OnePerseid meteor was classified as Na-poor, pointing out the possi-ble effect of cosmic-ray irradiation causing volatile depletion onHalley-type orbits.

The brightest and most detailed Perseid spectrum in our sam-ple can be seen in Fig. 13. The corresponding meteor was esti-mated to be of -9.5 mag, enabling us to observe more detailedspectral features compared to the majority of Perseid spectradominated by atmospheric emission. The characteristic Ca II ionH and K doublet at 393.4 nm and 396.9 nm is observed even byour system with low sensitivity in the near-UV. Other ion linestypical of the high-temperature component in bright and fast me-teors can be also seen: the Mg II line at 448.1 nm, Si II at 634.7nm, and Ca II at 854.2 nm. The Hα line (656.3 nm) of the Balmerseries, which can be associated with organic cometary content,is very bright in this spectrum. The spectrum showed very lowcontent of Fe and was classified as Fe-poor. The meteor started toradiate very high in the atmosphere (HB = 119.2 km) and disin-tegrated completely at around 82 km altitude, reflecting its veryfragile structure. All of the detected features correspond to thesoft cometary material of 109P/Swift-Tuttle.

Among other Halley-type showers, we have detected a pairof meteoroids from σ-Hydrids and Lyrids and three 49 An-dromedids showing relatively heterogeneous spectral properties.Some degree of heterogeneity in Fe/Mg and Na/Mg ratios (typi-cally up to 30% in ternary Na/Mg/Fe content) among meteoroidsfrom one parent object appears to be natural. This is suggestedfrom the wider spectral studies of Leonids, Geminids (Borovickaet al. 2005), Taurids (Matlovic et al. 2017), and Perseids (thiswork). Sodium enhancement was confirmed in one 49 Androme-did. One σ-Hydrid is classified as Na-poor as a result of its smallperihelion distance. Our sample also includes one Orionid me-teoroid with relatively high Fe intensity (Fig. 13). The meteorspectrum was observed during a meteor flare, which could causesome degree of Fe intensity overestimation due to saturationand optical thickness of the radiating plasma. Even consideringthese effects, the observed Fe intensity is unusually high for acometary body originating in comet 1P/Halley.

5.2. Ecliptical

Besides Taurids, ecliptical showers (Fig. 12) were most repre-sented in our sample by the κ-Cygnids (6 meteoroids) and α-Capricornids (4 meteoroids). κ-Cygnids are spectrally similar tothe Taurids, showing relatively similar Na content and highervariations of Fe intensity. The α-Capricornids show two distincttypes of spectra. We believe that these variations reflect struc-tural differences between the stream meteoroids. The two typesof α-Capricornids exhibited distinct light curves. The first typeshows spectra with increased Fe intensity observed during brightflares associated with sudden disruption. The second type showsNa enhancement and lower Fe content with smooth light curveswithout flares. The characteristic light curves with bright flares atthe end of trajectory were previously observed by Madiedo et al.(2014b), who also analyzed α-Capricornid emission spectra andconcluded that the meteoroids and their parent body are of non-chondritic composition. The spectrum modeling methodology

Article number, page 8 of 19

Page 9: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

Pavol Matlovic et al.: Spectral and orbital survey of medium-sized meteoroids

20

40

60

80

100

20

40

60

80

100

20 40 60 80

Fe

Mg

Na

F I ­ 15e

I ­ 2Mg I ­ 1Na

49 AndromedidsLyridsOrionidsPerseidssigma Hydrids

Halley type

(a)

20

40

60

80

100

20

40

60

80

100

20 40 60 80

Fe

Mg

Na

F I ­ 15e

I ­ 2Mg I ­ 1Na

alpha Capricornidseta Virginidskappa Cygnidsnorthern Tauridssouthern Taurids

Jupiter family / Ecliptical

(b)

Fig. 12: Spectral classification for identified meteors from established meteoroid streams. Separately plotted are meteoroid streamswith Halley-type orbits (a) and Jupiter-family type or ecliptical orbits (b).

Ca II

Mg IIFe I

Mg I

Fe I

Na I

N2

O I

Si II

N2

O I

H I

N2

N I

0

2000

4000

6000

8000

400 450 500 550 600 650 700 750 800Wavelength [nm]

Sig

nal [

arb.

u.]

M20160813_214954

O I

Ca II

Fe I Fe I

Mg I

Fe I

Fe I

Na I

O I

Si II

N2

O I

N I

0

200

400

600

400 450 500 550 600 650 700 750 800Wavelength [nm]

Sig

nal [

arb.

u.]

M20151024_003234O I

Fig. 13: Spectral profiles of a bright Perseid me-teor M20160813_214954 and a Orionid meteorM20151024_003234.

applied by Madiedo et al. (2014b) is however not suitable forthe low-resolution spectra and the presented abundances need tobe confirmed using higher-resolution data. Overall, the detectedlight curves and spectra of α-Capricornids suggest structural andcompositional heterogeneity of the meteoroid stream originatingfrom the inactive comet 169P/NEAT (Jenniskens & Vaubaillon2010).

20

40

60

80

100

20

40

60

80

100

20 40 60 80

Fe

Mg

Na

F I ­ 15e

I ­ 2Mg I ­ 1Na

alpha CapricornidsGeminidskappa CygnidsPerseidsSouthern delta AquariidsTauridszeta Cassiopeiids

Fig. 14: Mean spectral classification (mean Na I/Mg I/Fe I inten-sity ratios and standard error of the mean) of meteoroid streamswith at least three meteoroids in our sample.

The spectral differences in mean Na/Mg/Fe ratios betweenthe major meteoroid streams in our sample are displayed on Fig.14. This plot shows that while the spectral heterogeneity withinone stream can be quite substantial (Fig.12), the mean Na/Mg/Feratios can be used to distinguish between different streams. Foraccurate characterization of a mean spectral type of a meteoroid

Article number, page 9 of 19

Page 10: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

A&A proofs: manuscript no. AA_2019_36093_p

Table 1: Atmospheric and orbital properties of identified Perseid meteoroids. For references, see Tables 5 and 6 (available at theCDS).

Code Spec. type Mag αg δg vg HB HE q i ω TJ KB PE

M20150812_232102 Fe poor -7.1 46.17 57.90 61.51 126.80 86.88 0.965 114.55 155.45 - 6.00 D -5.82 IIIB± 1.5 0.05 0.03 0.58 0.18 0.11 0.001 0.32 0.57

M20150813_020206 Fe poor -4.2 47.99 58.03 56.34 110.87 82.60 0.937 111.13 145.38 1.03 6.75 C3 -5.11 II± 1.0 0.09 0.01 0.41 0.17 0.02 0.002 0.30 0.88

M20160812_005618 Fe poor -5.6 45.90 57.35 59.50 109.83 79.92 0.960 113.97 153.20 -0.24 6.93 C2 -5.07 II± 1.0 0.06 0.02 0.22 0.17 0.11 0.001 0.15 0.26

M20160813_214954 Fe poor -9.5 53.15 57.51 57.52 119.18 82.16 0.917 113.34 141.86 0.69 6.37 D -5.81 IIIB± 0.6 0.08 0.05 0.66 0.13 0.07 0.004 0.44 1.30

M20150810_013034 Na poor -3.4 46.10 60.72 56.16 107.64 90.28 0.925 106.69 144.38 0.37 7.00 C2 -5.54 IIIA± 1.3 0.07 0.02 0.79 0.18 0.14 0.004 0.60 1.41

M20150807_002948 normal -3.0 41.59 56.60 59.46 109.99 86.79 0.940 113.19 148.41 -0.41 6.92 C2 -5.10 II± 1.4 0.04 0.02 0.25 0.03 0.01 0.001 0.12 0.22

M20150808_232655 normal -5.1 42.10 58.66 59.46 119.90 75.59 0.953 111.05 151.90 - 6.30 D -4.69 II± 1.3 0.05 0.02 0.08 0.06 0.01 0.000 0.04 0.13

M20150811_005737 normal -5.0 48.45 56.63 58.16 115.95 83.84 0.919 113.81 142.88 0.40 6.48 D -5.28 IIIA± 1.1 0.04 0.02 0.12 0.04 0.01 0.001 0.07 0.27

M20150813_011318 normal -3.4 47.64 58.25 60.66 112.71 83.26 0.955 113.73 152.68 - 6.74 C3 -4.92 II± 0.7 0.07 0.03 1.38 0.31 0.03 0.003 0.72 1.10

M20150814_020823 normal -8.8 48.72 59.29 59.12 117.19 83.14 0.952 111.71 151.57 -0.40 6.37 D -5.90 IIIB± 1.3 0.04 0.11 0.66 0.17 0.16 0.002 0.46 0.80

M20160809_003127 normal -3.5 45.35 58.01 58.72 114.09 87.75 0.936 111.86 147.34 -0.20 6.63 C2 -5.30 IIIA± 0.8 0.06 0.03 0.37 0.23 0.04 0.002 0.24 0.59

M20160811_195200 normal -6.0 45.91 57.70 59.38 121.57 88.68 0.958 113.40 152.70 -0.27 6.30 D -5.61 IIIA± 0.8 0.35 0.27 1.23 1.18 0.56 0.003 0.62 1.50

M20160811_230132 normal -5.0 47.66 56.70 59.02 109.51 82.46 0.946 114.68 149.24 0.11 6.99 C2 -5.06 II± 0.5 0.03 0.09 0.56 0.17 0.10 0.002 0.29 0.87

M20160812_002835 normal -4.6 47.58 57.70 57.96 112.84 80.84 0.944 112.65 148.30 0.38 6.71 C2 -5.04 II± 0.8 0.10 0.06 0.40 0.33 0.22 0.002 0.27 0.64

M20160812_011115 normal -1.8 44.82 58.31 62.36 101.71 85.58 0.972 114.34 157.80 - 7.54 A -4.72 II± 0.6 0.04 0.03 0.67 0.07 0.42 0.001 0.36 0.56

M20160812_011418 normal -5.3 45.95 57.80 58.75 107.14 82.26 0.958 112.91 152.27 0.00 7.09 C2 -5.19 II± 0.7 0.10 0.37 0.63 0.55 0.42 0.002 0.48 1.05

M20160812_014157 normal -7.2 48.65 59.96 57.30 101.86 75.08 0.936 109.32 147.05 0.20 7.42 A -5.04 II± 0.8 0.05 0.33 0.82 0.83 0.63 0.004 0.60 1.64

M20150811_012547 Na enhanced -3.0 43.22 56.75 58.38 103.85 80.97 0.960 113.35 152.31 0.28 7.31 A -4.68 II± 1.0 0.52 0.12 0.78 0.34 0.13 0.004 0.54 1.45

M20150812_011600 normal -2.5 43.62 56.53 59.50 114.15 86.06 0.968 114.62 155.20 -0.12 6.62 C2 -5.02 II± 1.0 0.05 0.01 0.56 0.20 0.05 0.001 0.34 0.63

stream, sufficient statistical sample is required (over 10 meteors),which is only satisfied for Perseids and Taurids in our sample.

Figure 14 also shows the spectral similarity of Taurids andκ-Cygnids (and potentially also α-Capricornids). This might im-ply that the two streams have parent objects of similar compo-sition. Similarly to Taurids, κ-Cygnids are thought to be rem-nants of a previous large comet disruption. The largest remainingobject from the break-up associated with κ-Cygnids is the dor-mant 2008 ED69 (Jenniskens & Vaubaillon 2008). The spectrumof this body is not known, but the fragile structure and chon-dritic abundances found in one κ-Cygnid by Trigo-Rodriguezet al. (2009) could suggest a C- or D-type body. On the otherhand, comet 2P/Encke associated with the major stream of Tau-rids is Xe-type but has a spectrum resembling primitive asteroids(Tubiana et al. 2015). The link between Taurids and primitivecarbonaceous materials was also supported by the two associatedmeteorite falls of Maribo and Sutter’s Mill (Tubiana et al. 2015).Overall, this could imply a similar primitive (C- or D-type) com-position of the parent objects of κ-Cygnids and Taurids.

5.3. Unconfirmed showers

The showers in the IAU MDC working list are in the pro-cess of being confirmed as real individual meteoroid streams.A large survey by Jenniskens et al. (2016b) recently showed thatmany of these streams show significant annual activity. Here,we only report on mainly individual meteor spectra associatedwith working list showers, which if confirmed could providethe first spectral data for these streams. The identified Halley-type and Jupiter-family/ecliptical working list shower meteorsare displayed in Fig. 15.

The spectra of Halley-type shower meteors are mostly nor-mal type, with no notable differences from other major streammeteoroids. One of the three ζ-Cassiopeiids showed slight Nadepletion and was classified as Na-poor. Overall, their spectralcharacteristics are similar to Perseids, though on average theyexhibit a slight increase in Fe/Na. More distinct spectral classeswere detected among Jupiter-family and ecliptical shower mete-ors (Fig. 15). Specifically, one Na-enhanced meteoroid was as-signed to the ν-Draconids stream, which was reported as a du-

Article number, page 10 of 19

Page 11: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

Pavol Matlovic et al.: Spectral and orbital survey of medium-sized meteoroids

20

40

60

80

100

20

40

60

80

100

20 40 60 80

Fe

Mg

Na

F I ­ 15e

I ­ 2Mg I ­ 1Na

10 Canum Venaticids12 Bootidsf Herculidsmu Leonidsmu Pegasidsomega Ursae Majoridsr Lyridsupsilon Cetidszeta Bootidszeta Cassiopeiids

Halley type

(a)

20

40

60

80

100

20

40

60

80

100

20 40 60 80

Fe

Mg

Na

F I ­ 15 e

I ­ 2Mg I ­ 1Na

64 Draconidschi Libridsgamma Libridsmu Virginids nu Dr aconids Dec. rho Geminidszeta1 Cancrids

Jupiter family / Ecliptical

(b)

Fig. 15: Spectral classification for identified meteors from unconfirmed (working list) meteoroid streams. Separately plotted aremeteoroid streams with Halley-type orbits (a) and Jupiter-family type or ecliptical orbits (b).

plicate of the established August Draconids (Jenniskens et al.2016b). The shower is active in August/September with charac-teristically low-velocity meteors (vg = 20.1 km s-1)). The parentbody of the stream is not known. As a preliminary estimate, wesuspect that the detected Na-enhancement is mainly caused bythe low speed of the meteor and does not reflect atypical compo-sition within the stream. This is supported by the detected chon-dritic Fe/Mg ratio and no identified anomalous lines.

One December ρ−Geminid (DRG) meteoroid was detectedwith a Na-free spectrum, similarly to the typical spectra(Borovicka et al. 2005) of the major Geminid stream of 3200Phaeton. Jenniskens et al. (2016a) identified DRG meteoroidsas members of the Geminid complex. It is assumed that dustparticles released from Phaethon, 2005 UD, and potentially an-other yet-to-be discovered parent body contribute to the streamswithin the Geminid complex (Jenniskens et al. 2016a). Decem-ber ρ−Geminid meteoroids have been identified as unusually fastGeminids, which is also confirmed in our sample (vg = 40.5km s-1). Jenniskens et al. (2016a) speculated that the observedunusual deceleration of these meteoroids could be linked withatypical physical properties and low density. Our sample showsspectrum depleted in Na as a result of strong solar radiation, andhigh material strength (type ast/I) of the meteoroid, characteris-tic for ordinary chondrites. Both spectrally and physically, thismeteoroid fits well with other Geminids from the main streamand asteroidal parent body, but the reason behind the high speedof this branch remains unclear.

6. Structure

The beginning and terminal (end) heights of meteor trajectoriescan reveal structural properties of the incoming meteoroid. Theyare used as a basis of the empirical material strength parame-ters KB and PE introduced by Ceplecha (1968) and Ceplecha &McCrosky (1976). In general, fragile particles tend to crumbleunder lower pressures resulting in higher beginning heights of

70

80

90

100

110

120

130

10 20 30 40 50 60 70Meteor speed [km/s]

Beg

inni

ng h

eigh

t [km

]

Fe poorFe richIronsNa enhancedNa freeNa poorNa richNormal

Fig. 16: Meteor beginning height as a function of entry speedfor different spectral types of meteors. The solid line representsa quadratic fit to the running average of our data and standarderror of the mean. The dashed line represents the fit of the datafor fainter meteors from Vojácek et al. (2019). Orange and redlines show a theoretical model for cometary and asteroidal me-teors, respectively, simulated for systems of comparable sensi-tivity (CAMS-type) by Vida et al. (2018).

the luminous trajectory. Likewise, only stronger materials canwithstand the exerted pressure and reach lower parts of our at-mosphere, and only the most robust bodies will eventually fall tothe Earth’s surface as meteorites. It is observed that dependingon the entry velocity and meteoroid mass, the ablation begins atgreater heights for cometary bodies compared to asteroidal ones(Ceplecha & McCrosky 1976; Koten et al. 2004).

The meteor beginning height is dependent on meteor speedas displayed in Fig. 16. A similar dependency was observedamong fainter meteors by Vojácek et al. (2019), but shifted lower

Article number, page 11 of 19

Page 12: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

A&A proofs: manuscript no. AA_2019_36093_p

30

40

50

60

70

80

90

100

60 70 80 90 100 110 120 130Beginning height [km]

Term

inal

hei

ght [

km]

Fe poorFe richIronsNa enhancedNa freeNa poorNa richNormal

Fig. 17: Meteor beginning height vs. meteor terminal height plotfor different spectral types of meteors.

by approximately 8 km lower due to the different sensitivity ofthe system. Our results are also consistent with the simulationsof Vida et al. (2018), which suggest that two branches of the plotcould approximately recognize asteroidal and cometary bod-ies. The lowest beginning heights are observed among Na-richand Fe-rich meteoroids (typically 73 km < HB < 83 km com-pared to most common meteor beginning heights in the range90 km < HB < 115 km). Low beginning heights are also ob-served for most Na-enhanced meteoroids. In all three groups, thelow beginning heights are explained by generally high materialstrength associated with chondritic-asteroidal bodies. Figure 16also shows that Na-poor and Na-free meteoroids have lower be-ginning heights than normal-type meteors of similar speed. Onthe other hand, most Fe-poor meteors typically show cometaryproperties with higher beginning heights.

A similar functionality is found for meteor terminal heights.The link between beginning and terminal heights is displayedin Fig. 17. Although terminal heights are more strongly influ-enced by meteoroid mass, Koten et al. (2004) have shown thatbeginning heights are also mass dependent. The range of mete-oroid beginning heights for different spectral types is well dis-tinguished in Fig. 18: 110 - 127 km for Fe-poor, 78 - 123 km fornormal type, 85 - 119 km for Na-poor, 89 - 97 km for Na-free,81 - 112 km for Na-enhanced, 78 - 95 km for Fe-rich, and 73- 80 km for Na-rich. This plot refers to the characteristic mate-rial strengths and structural heterogeneity among each spectralclass. The widest range is predictably observed among normal-type spectra, which include both cometary and chondritic me-teoroids. Still, the majority of normal-type meteoroids have be-ginning heights between 100 and 120 km, characteristic of morefragile, cometary bodies. This preference might be affected bythe higher number of shower meteors in our sample. Signifi-cant heterogeneity is also found among Na-enhanced meteors,this time dominated by stronger meteoroids with lower begin-ning heights. This distinction also shows the inclusion of bothcometary and asteroidal meteoroids in this spectral class, as sup-ported by the determined orbits.

70

80

90

100

110

120

130

−1.5 −1.0 −0.5 0.0 0.5 1.0 1.5 2.0Na/Mg (log10)

Beg

inni

ng h

eigh

t [km

]

Fe poorFe richIronsNa enhancedNa freeNa poorNa richNormal

Fig. 18: Dependence of Na/Mg intensity ratio on meteor begin-ning height for different spectral types of meteors. The verticalcolumns represent the range of detected beginning heights foreach spectral class positioned at mean Na/Mg value.

40

60

80

100

−0.5 0.0 0.5 1.0Fe/Mg (log10)

Term

inal

hei

ght [

km]

Fe poorFe richIronsNa enhancedNa freeNa poorNa richNormal

Fig. 19: Dependence of Fe/Mg intensity ratio on meteor terminalheight for different spectral types of meteors.

To some degree, the differences in Fe/Mg ratio can be used todistinguish between cometary and asteroidal bodies. Meteoroidswith low iron content exhibit higher terminal heights, charac-teristic for fragile meteoroids (Fig. 19). The average terminalheight for meteors with Fe/Mg < 0.7 is HE = 79 km, whilefor meteors with Fe/Mg > 1.0 it is HE = 69 km. The depen-dency is not completely straightforward due to the strong ef-fect of entry speed and meteoroid mass. It also only concernsstony meteoroids, since the ablation of iron-rich bodies is atyp-ical and cannot be described by the standard single-body the-ory. Capek & Borovicka (2017) have shown that iron meteoroidshave typically relatively short light curves with low beginningheights but terminal heights usually above 70 km. Similarly, Fig.19 shows that most Fe-rich meteoroids and the only iron mete-oroid in our sample had intermediate terminal heights between80 and 60 km. The lowest terminal height (HE = 33.0 km) wasdetected in the brightest meteor in our sample, the Na-poor me-teor M20170227_023123 of estimated -11.2 mag.

The corresponding material strength classification based onKB and PE parameters for all meteoroids observed by multiple

Article number, page 12 of 19

Page 13: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

Pavol Matlovic et al.: Spectral and orbital survey of medium-sized meteoroids

stations is displayed in Fig. 20. The corresponding values can befound in Table 6 (available at the CDS). Since reliable meteoroidmasses were not determined, the dependence of beginning heighton original meteoroid mass suggested by Koten et al. (2004)cannot be clearly evaluated. Still, no clear preference of meteormagnitude is observed in the KB classification. The PE distribu-tion shows that mostly bright fireballs are found among the mostfragile IIIB meteoroids, including seven fireballs brighter than-8 mag. The meteor magnitude does not always fully reflect onoriginal meteoroid mass. For example, the strong Na-rich me-teoroids are of moderate magnitude, even though we estimatedrelatively high photometric masses.

Figure 20 shows that Na-rich, Na-free, and Na-poor mete-oroids are on average composed of the strongest material. Nu-merous strong, chondritic materials are also identified amongnormal-type meteors, though none of them are found in the ordi-nary chondrites group (ast) in the KB classification. High mate-rial strength KB is also detected in Fe-rich meteoroids, thoughthe PE values are scattered and generally low. This behaviorcould reflect the enhanced iron content, as the ablation of ironmeteoroids is atypical with relatively short light curves andmoderate terminal heights. The same pattern is observed in theonly iron meteoroid in our sample which exhibits high materialstrength at the beginning of ablation (ast), but dissipates quiteearly (IIIA). The lowest material strength is on average found inFe-poor meteoroids, similarly to the previous results for smallermeteoroids (Vojácek et al. 2015).

Our results also demonstrate the correlation between the de-termined material properties and meteor orbits (Fig. 20). Mete-oroids originating in the main asteroid belt or from NEO orbits(TJ > 3) are on average of higher material strength, most oftencharacteristic for carbonaceous and ordinary chondrites. A fewfragile meteoroids were also found on asteroidal orbits, mainlyincluding shower meteors of the Taurids and α-Capricornids.This population represents fragments of short-period comets,which are deficient compared to asteroidal bodies in this region,but produce a larger amount of dust particles. In case of the Tau-rids, there are hints that the stream also contains debris from as-teroids on similar orbits, though no chondritic meteoroids wereidentified in the sample from our previous study (Matlovic et al.2017).

The range of materials detected on the Jupiter-family orbitsis wider and includes both asteroidal and cometary meteoroids(Fig. 20). On average, the majority of the meteoroids in thisregion are defined by the cometary and carbonaceous materialstrengths (types C-A/IIIA-II). Several meteoroids with materialproperties similar to ordinary chondrites are also present, in largepart represented by Na-poor and Na-free spectra. In this case,the hardening of material during close approaches to the Sun,associated with the release of volatiles, may be responsible forthe detected high material strengths (see section 7.1). The KBvalues show no meteoroids composed of the softest cometarymaterial (type D) present on Jupiter-family orbits. The presenceof the most fragile cometary material in the Jupiter-family zoneis however possible, as four samples were identified as PE typeIIIB. Two of these meteoroids are members of the κ-Cygnids andtwo are the α-Capricornids.

Our results also suggest that meteoroids from Halley-type or-bits do not contain ordinary chondrites (type ast/I). The only ex-ception is one Na-poor meteoroid. Another normal-type meteorM20140726_000917 on a Halley-type orbit was seen to havetypically cometary material strength at the ablation beginning(C), but unusually high PE material strength. The reason behindthis discrepancy is not clear. The meteoroid could represent a

Fig. 20: Material strength classification of all meteoroids ob-served by multiple stations based on the KB (upper) and PE(lower) parameter as a function of the Tisserand parameter.D/IIIB type represents the most fragile cometary material,C/IIIA typical cometary material, B denser cometary material,A/II material similar to carbonaceous bodies, and ast/I materialsimilar to ordinary chondrites. Sizes of meteoroid marks reflectrelative meteor magnitudes. Color coding differentiates betweenthe determined meteor spectral classes.

denser cometary material, possibly formed near a cometary nu-cleus. Overall, the meteoroids on Halley-type orbits are mainlyof fragile structure, including several meteoroids of the softestcometary material (D/IIIB). The most fragile structure was typi-cally found among brighter fireballs and included several streammeteoroids, including the Perseids, Orionid, and ζ-Cassiopeidamong Halley-type streams, and κ-Cygnids and α-Capricornidsamong Jupiter-family streams. This material class D/IIIB wasoriginally defined to include the brittle Giacobinids (now Dra-conids). Such soft material is apparently not unusual amongother meteoroid streams.

Even though Perseid meteoroids were on average of char-acteristically fragile material strength, Table 1 shows that some

Article number, page 13 of 19

Page 14: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

A&A proofs: manuscript no. AA_2019_36093_p

meteoroids showed higher strength comparable to carbonaceousbodies. This is particularly apparent from the PE values, whereover half of our samples were identified as type II. Similarly tothe spectra, we observe high heterogeneity of material strengthswithin Perseids. Partial results for different showers suggest thatthis is also common among other meteoroid streams. The Per-seid spectral and material properties show some degree of cor-relation. In general, Fe-poor Perseids are more fragile and theNa-enhanced sample is among the strongest (A/II). Unlike Na-depleted meteoroids affected by strong solar radiation, the Na-poor Perseid was of low material strength.

The material strength classification of Halley-type mete-oroids (Fig. 20) includes numerous A/II type bodies withstrengths similar to carbonaceous materials. The abundance ofthese meteors is particularly high in the PE distribution (Fig.20). This does not imply that all meteoroids characterized astype II are of carbonaceous composition; this group likely alsoincludes dense cometary bodies and fractured chondrites. Ce-plecha (1988) noted that this group includes both cometary andasteroidal bodies, as observed in our data. The dominance ofcometary materials on Halley-type orbits is apparent in our data.Recently, Vojácek et al. (2019) reported the existence of iron me-teoroids on Halley-type orbits, showing that the evolution of theearly solar system could result in a more complicated distribu-tion of materials. The complex mixing and transport of materialswas previously revealed from the samples of comet 81P/Wild 2collected by the Stardust mission (Brownlee et al. 2012).

7. Spectral groups

7.1. Na-poor and Na-free

The Na line is one of the most interesting spectral features inmeteors. In general, it corresponds to the volatile content in me-teoroids. The dependence is however not straightforward, as theNa line intensity in meteors is also influenced by physical con-ditions, particularly meteor speed. Still, while real Na enhance-ment can be difficult to reveal, the depletion of Na is directlylinked with the loss of volatiles within the meteoroid composi-tion. The detected Na depletion can reveal the thermal evolutionof the meteoroid. The main processes of Na loss in meteoroidsconcern space weathering caused by thermal desorption duringclose perihelion approaches and cosmic-ray irradiation in theOort cloud for long-period orbits (Borovicka et al. 2005). Ironmeteoroids have naturally low Na content.

We identified 18 Na-poor and 5 Na-free meteors in our sam-ple, and orbital data is available for 12 and 4 of these meteors,respectively. In terms of spectra, the main difference between thetwo classes is that the Na line can still be reliably measured inNa-poor meteors. As noted earlier, the detected fraction of Na-depleted meteors is significantly lower compared to the resultsof studies focused on fainter meteors (Borovicka et al. 2005; Vo-jácek et al. 2015, 2019). We argue that this effect reflects the in-creased preservation of volatiles in larger meteoroids. The spaceweathering processes more significantly affect smaller bodies.This corresponds to the fact that most Na-poor and Na-free spec-tra were observed in fainter meteors within our sample. Previ-ously, Borovicka et al. (2005) noted no correlation between theNa content and meteoroid mass within the size range of theirsample.

Of the processes causing Na depletion in meteoroids, thermaldesorption was identified as the most common. The dependencyof Na/Mg ratio on perihelion distance (q; Fig. 21) shows that allfour measured Na-free meteors have q < 0.2 au. This correlates

−1

0

1

2

0.0 0.2 0.4 0.6 0.8 1.0Perihelion distance [au]

Na/

Mg

(log 1

0)

Fe poorFe richIronsNa enhancedNa freeNa poorNa richNormal

Fig. 21: Perihelion distance vs. measured Na/Mg intensity ratiofor meteoroids of different spectral classes.

with the results of Borovicka et al. (2005) suggesting that Naloss by thermal desorption occurs at q ≤ 0.2 au. According toKasuga et al. (2006), sodium abundance in meteoroids does notdepend on their perihelion distances at q ≥ 0.14 au. This limit issatisfied by three Na-free meteors. Furthermore, Na loss is likelycorrelated with the dynamic age of a meteoroid. Younger mete-oroids have suffered fewer passages in the vicinity of the Sunand can retain more Na. The heating during close perihelion ap-proaches causes loss of volatiles and consequent compaction ofthe meteoroid material. This effect might be related to the aque-ous alteration and is demonstrated by the generally high mate-rial strength of Na-free, and to some degree Na-poor meteoroids(Fig. 16 and 20). As seen in Fig. 20, the KB and PE classificationplaces Na-free and Na-poor among the strongest spectral classes.

In general, meteors classified as Na-poor have small or inter-mediate perihelion distances (q ≤ 0.37 au) or are on long-periodcometary orbits. The process of sodium loss on orbits with 0.2< q < 0.4 au is not clear. The thermal effects from solar radia-tion should not be as significant here (Kasuga et al. 2006). It ispossible that some degree of thermal loss in combination withthe aging effect is still occurring. We also observed a cluster ofNa-poor meteors on Halley-type orbits with 0.9 < q < 1.0 au(Fig. 21). The Na-depleted material in these meteoroids mightcome from cometary crust affected by long-term cosmic ray ir-radiation, as previously suggested by Borovicka et al. (2005).Our results suggest that this effect is not as significant as solarheating, as no Na-free meteors on such orbits were detected. Theatmospheric and orbital properties of Na-free and Na-poor me-teoroids are given in Tables 2 and 3.

The most atypical meteor classified as Na-poor isM20170227_023123 (Fig. 22). This meteor is by far the bright-est of all Na-depleted bodies in our sample, with an estimated-11.2 mag. Among the Na-poor class, it is positioned in the up-permost part of the ternary diagram (Fig. 6), reflecting chondriticvalues of the Fe/Mg ratio. The meteor had an unusual monochro-matic light curve with early release of Mg and delayed emissionof Fe and Na lines (Fig. 22). This behavior is different from themore commonly observed early release of volatile Na. The spec-trum of M20170227_023123 could reflect its larger size with Nadepleted from its outer layers. A similar monochromatic lightcurve was observed in one atypical Geminid by Vojácek et al.(2019). The late release of Na was explained by different grainsizes in the two stages of erosion of the meteoroid. Faster deple-

Article number, page 14 of 19

Page 15: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

Pavol Matlovic et al.: Spectral and orbital survey of medium-sized meteoroids

Ca IFe I

Mg I

Mg II Fe I

Mg I

Fe I

Fe I

Mg I

Na I

O I

FeO

O I N I

O I

0

1000

2000

400 500 600 700 800Wavelength [nm]

Sig

nal [

arb.

u.]

M20170227_023123

Fig. 22: Spectrum of a bright Na-poor meteorM20170227_023123 with identified major lines (upper).The monochromatic light curve (lower) shows early emission ofMg, which also remains after Na and Fe lines disappear. Here,the Fe I monochromatic light curve is only represented by theintensity of the 526.9 nm line. One frame corresponds to a timeperiod of 1/12 s.

tion of Na is expected for smaller grains contained in the part ofthe meteoroid which fragmented first.

This meteoroid was identified as a member of the η-Virginidstream. The orbital data from different authors for the streamshow a wide range of mean perihelion distances (0.23 ≤ q ≤ 0.46au)4. Similarly, two η-Virginids were identified in our sample,M20170227_023123 with q = 0.33 au and M20170314_194345with q = 0.47 au. The two η-Virginid have also shown signifi-cant spectral differences (Fig. 24), which could imply a hetero-geneous composition within the stream. The parent object of η-Virginids is not confirmed, but there are hints that the streamcould be linked with the short-period comet D/1766 G1.

Several other meteoroid streams were identified as a sourceof Na-free and Na-poor bodies. Unsurprisingly, the associatedstreams have typically small perihelion distances. One Na-freeand one Na-poor meteor were found as members of the Southernδ-Aquarids (q = 0.07 au). The δ-Aquarids are a Sun-approachingstream presumably formed from the breakup of a larger body,which also formed the Marsden and Kracht sungrazing comets(Jenniskens 2006). As probably the largest remaining fragmentof this breakup, 96P/Machholz is assumed to be responsible formost of the dust production to the stream. While the very small qorbits are a natural cause for Na depleted composition, we havealso detected one δ-Aquarid meteor with a normal-type spec-trum. This spectrum still shows a somewhat lower intensity ofNa compared to most normal-type meteors (Fig. 24). The differ-

4 A collection of mean shower orbits determined bydifferent authors can be found in the IAU MDC:https://www.ta3.sk/IAUC22DB/MDC2007/ (Jopek & Kanuchová2017)

Ca IFe I

Fe I

Mg I

Fe I

O I

N2

O I

N I

O I

0

100

200

300

400 500 600 700 800Wavelength [nm]

Sig

nal [

arb.

u.]

M20140726_001002

Ca IFe I

Mg IFe I

Mg I

Fe I

Na I

O I

N2

O I

N I O I

0

200

400

400 500 600 700 800Wavelength [nm]

Sig

nal [

arb.

u.]

M20140803_010827

Fig. 23: Spectra of two δ-Aquarid meteors. The difference be-tween Na-free (upper) and normal-type (lower) spectra showsthe aging effect among meteoroids in a Sun-approaching stream.

ence between the Na-free and normal δ-Aquarid spectra is shownin Fig. 23. The normal-type sample was a meteor of -5.0 magcompared to -3.1 mag for the Na-free meteor. This could indi-cate that Na is better preserved in the larger meteoroid, but thesize difference is probably not sufficient to explain the spectra.We suggest that the difference between the spectra demonstratesthe aging effect of meteoroids on orbits with small q. The differ-ence also implies that the meteoroids are being actively suppliedto the δ-Aquarid stream. We previously pointed out the diversityof Na/Mg ratios in the two δ-Aquarids in Rudawska et al. (2016).

In addition to δ-Aquarids, Borovicka et al. (2005) identi-fied Geminids (q = 0.14 au) and α-Monocerotids (q = 0.19 au)as meteoroid streams containing Na-free and Na-poor meteors.Surprisingly few bright Geminids were collected during the rou-tine observations of the AMOS-Spec in 2013-2017. The obser-vation campaigns for Geminids were in some cases hamperedby bad weather, often leading to only single-station data for po-tential Geminid spectra. Still, several Na-poor spectra were de-tected near the predicted maximum of the shower activity, asdisplayed in Fig. 24. Borovicka et al. (2005) noted a wide spreadof Na line intensity in Geminid spectra, also explained by theaging effect. The positions of Geminids identified in Fig. 24 cor-relate well with the results of Borovicka et al. (2005). It appearsthat Geminids are much better represented in spectral surveys offainter meteors, suggesting that the stream could include fewerlarger centimeter- to meter-sized debris. In the future, we intendto carry out a more extensive study of the Geminid compositionbased on high-resolution spectra from global AMOS-Spec sta-tions.

Furthermore, one Na-free meteor (M20151227_035406) wasassigned to the December ρ-Geminids (Jenniskens et al. 2016a).The ρ-Geminids are a IAU MDC working-list shower character-ized by higher velocities (vg = 39.5 km s-1) and smaller perihe-lion distances (q = 0.11 au) compared to the Geminids. One σ-Hydrid meteor (q = 0.22 au) was found to have a Na-poor spec-

Article number, page 15 of 19

Page 16: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

A&A proofs: manuscript no. AA_2019_36093_p

Table 2: Atmospheric and orbital properties of detected Na-free meteoroids observed by multiple stations. For references, see Tables5 and 6 (available at the CDS).

Code Shower αg δg vg HB HE a e q Q i ω Ω TJ

M20140726_001002 SDA 336.82 -15.22 41.02 96.81 81.49 2.39 0.97 0.062 4.73 25.80 154.51 302.77 2.45± 0.01 0.04 0.05 0.08 0.05 0.01 0.00 0.001 0.02 0.07 0.10

M20150710_232522 spo 315.16 -4.01 39.07 96.28 81.19 2.13 0.94 0.118 4.15 36.84 325.00 108.17 2.78± 0.02 0.05 0.17 0.06 0.02 0.04 0.00 0.001 0.09 0.43 0.04

M20150829_015634 spo 354.71 -1.52 33.11 92.50 67.77 2.15 0.90 0.207 4.09 1.27 312.43 155.20 2.97± 0.03 0.03 0.09 0.08 0.14 0.02 0.00 0.001 0.04 0.04 0.09

M20151227_035406 DRG 128.32 25.39 40.50 89.49 67.09 2.01 0.96 0.072 3.94 23.35 332.60 274.84 2.89± 0.10 0.07 0.75 0.18 0.16 0.25 0.00 0.004 0.50 1.35 0.43

Table 3: Atmospheric and orbital properties of detected Na-poor meteoroids observed by multiple stations. For references, seeTables 5 and 6 (available at the CDS).

Code Shower αg δg vg HB HE a e q Q i ω Ω TJ

M20131203_050007 DKD 187.13 72.22 43.72 100.37 88.80 39.17 0.98 0.928 77.40 71.77 208.27 251.00 0.50± 0.67 0.05 0.67 0.24 0.21 - 0.04 0.002 - 0.65 0.53

M20150308_020714 XHE 256.57 42.94 37.13 86.54 61.15 2.08 0.53 0.986 3.17 66.39 191.28 346.97 2.94± 0.04 0.03 0.07 0.05 0.03 0.02 0.00 0.000 0.04 0.08 0.06

M20150409_203107 spo 221.65 9.31 37.82 87.65 77.06 11.42 0.97 0.364 22.47 37.95 287.09 19.46 1.04± 0.39 0.08 0.69 0.47 0.64 - 0.02 0.004 - 0.91 0.80

M20150518_234943 spo 292.75 67.24 26.75 91.83 56.33 2.66 0.62 1.001 4.31 44.76 166.52 57.50 2.75± 0.11 0.04 0.31 0.14 0.09 0.12 0.02 0.000 0.24 0.36 0.22

M20150803_202646 spo 301.25 61.37 33.86 98.77 82.09 11.62 0.92 0.988 22.26 54.28 198.94 130.97 1.15± 0.06 0.04 0.26 0.12 0.10 3.21 0.02 0.000 6.39 0.25 0.17

M20150810_013034 PER 46.10 60.72 56.16 107.64 90.28 7.44 0.88 0.925 13.96 106.69 144.38 136.92 0.37± 0.07 0.02 0.79 0.18 0.14 - 0.06 0.004 - 0.60 1.41

M20151210_232747 HYD 125.59 3.44 58.03 105.81 82.76 19.25 0.99 0.209 38.30 127.10 125.79 78.36 -0.07± 0.03 0.09 0.24 0.15 0.15 - 0.00 0.004 - 0.35 0.72

M20160122_195954 ACZ 137.80 17.57 31.60 85.72 60.52 2.71 0.89 0.311 5.12 1.74 297.07 302.06 2.59± 0.02 0.02 0.05 0.05 0.02 0.02 0.00 0.001 0.04 0.02 0.04

M20160727_013114 ZCS 22.97 57.99 54.66 118.30 90.91 6.06 0.84 0.964 11.15 103.22 152.62 124.23 0.59± 0.09 0.03 0.88 0.20 0.02 8.92 0.07 0.003 - 0.68 1.25

M20160730_012149 SDA 340.16 -15.85 41.56 97.45 80.97 2.91 0.98 0.069 5.75 27.63 152.52 307.10 2.08± 0.04 0.07 0.10 0.06 0.07 0.02 0.00 0.000 0.03 0.25 0.08

M20160902_230214 spo 3.11 -18.46 30.20 94.83 56.38 2.35 0.85 0.364 4.33 22.83 113.29 340.66 2.88± 0.03 0.03 0.03 0.03 0.03 0.01 0.00 0.000 0.01 0.07 0.02

M20170227_023123 EVI 173.25 5.12 29.27 91.44 33.04 1.94 0.83 0.330 3.54 2.53 298.08 338.46 3.37± 0.16 0.07 0.06 0.06 0.02 0.03 0.00 0.002 0.06 0.03 0.37

Table 4: Atmospheric and orbital properties of detected iron and Fe-rich meteoroids observed by multiple stations. For references,see Tables 5 and 6 (available at the CDS).

Code Type αg δg vg HB HE a e q Q i ω Ω TJ

M20150508_002949 Fe-rich 63.67 66.17 14.35 79.42 64.19 2.32 0.60 0.921 3.72 16.15 140.16 46.91 3.27± 3.53 0.73 0.17 0.42 0.36 0.16 0.02 0.007 0.31 0.15 2.10

M20160416_211518 Fe-rich 113.31 35.55 8.53 82.58 37.42 2.71 0.63 1.002 4.43 3.04 173.82 27.06 3.04± 0.68 0.46 0.47 0.47 0.15 0.31 0.04 0.000 0.63 0.25 0.35

M20160623_210113 Fe-rich 257.53 55.35 26.08 94.80 79.49 - 0.99 1.005 - 37.89 192.46 92.57 1.03± 0.13 0.13 0.20 0.12 0.10 - 0.01 0.000 - 0.21 0.13

M20170301_201252 Fe-rich 114.83 21.81 7.75 77.94 61.33 1.84 0.47 0.964 2.71 0.07 203.36 341.43 3.88± 0.57 0.76 0.22 0.30 0.23 0.04 0.01 0.001 0.09 0.11 0.41

M20160415_002243 iron 231.60 11.20 27.24 84.43 68.89 1.34 0.68 0.423 2.25 31.78 295.19 25.25 4.53± 0.08 0.08 0.14 0.27 0.17 0.01 0.00 0.001 0.02 0.24 0.18

Article number, page 16 of 19

Page 17: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

Pavol Matlovic et al.: Spectral and orbital survey of medium-sized meteoroids

20

40

60

80

100

20

40

60

80

100

20 40 60 80 100

Fe

Mg

Na

Fe

Mg Na

eta VirginidsGeminidskappa Draconidsrho Geminidssigma HydridsSouthern delta Aquariidsxi Herculids

Fig. 24: Selected stream meteoroids with detected Na depletion.The displayed Geminids were estimated from single-station databased on radiant position and solar longitude.

trum. Another σ-Hydrid was identified as normal-type, thoughwith a lower-than-average Na/Mg ratio (Fig. 24). Furthermore,the κ-Draconids, ξ-Herculids, and η-Virginids were also found toproduce Na-depleted meteoroids, each with one identified spec-trum. Overall, our data on Na-poor meteors suggest that somedegree of Na depletion is present also on orbits with intermedi-ate perihelion distances 0.2 < q < 0.4 au, probably also relatedto the effects of solar radiation.

7.2. Fe-rich and irons

As noted earlier, only one pure iron meteoroid was identified inour sample. The overall low abundance of iron meteoroids in oursample compared to surveys of smaller meteoroids probably re-lates to the typical sizes of iron–nickel grains. Iron meteoroidsare generally found on asteroidal orbits with a likely origin in themain asteroid belt. However, it appears that the main belt mightnot be the only source of iron meteors. One iron body was foundby Vojácek et al. (2015) on a borderline Jupiter-family orbit, andVojácek et al. (2019) even pointed out the surprising existence ofiron meteoroids on Halley-type orbits. Common features are ob-served among irons, including similar trajectory heights, lengths,and light curves, which cannot be explained by simple single-body ablation theory.

The iron meteor M20160415_002243 (Fig. 25) from oursample shows orbital and atmospheric parameters typical for anasteroidal body (Table 4). The speed of the meteor (vi = 29.4km s-1) is higher than for most iron meteors (typically below 20km s-1). A similar speed has however already been detected inone meteor by Borovicka et al. (2005) and an even higher speedwas found among the iron meteors on Halley-type orbits by Vo-jácek et al. (2019).

We identified two types of meteoroids observed as Fe-richmeteors. The majority of the increase in Fe/Mg ratio can becaused by the enhanced presence of iron–nickel or iron sulfide

Fe I Fe I

Fe I

Fe I

Fe I

Fe I

FeO

Fe I

N I

O I

0

200

400

600

400 500 600 700 800Wavelength [nm]

Sig

nal [

arb.

u.]

M20160415_002243

Fe IFe I

Fe I

Fe I

Mg I

Fe I

Fe I

Ca I

Na I

FeO

Ca I

FeO

0

500

1000

1500

400 500 600 700Wavelength [nm]

Sig

nal [

arb.

u.]

M20160402_020222

Fig. 25: Spectral profile of an iron meteor M20160415_002243(upper) and a Fe-rich meteor M20160402_020222 (lower).

grains in smaller meteoroids. The most probable composition ofthese meteoroids is similar to H-type chondrites, with detectedFe enhancement either caused by heterogeneous content of irongrains or unresolved overestimation of Fe intensity due to the op-tical thickness of the radiating plasma. The effects of saturationand optical thickness were examined in each case. The spectrumof these meteors is often obtained from a few frames during sud-den meteoroid disruption (exhibited as a meteor flare). One suchexample is meteor M20150508_002949 which showed a lightcurve with numerous continuous flares, suggesting a fracturedstructure of a relatively strong material. A similar light curveand spectral type was detected in meteor M20170301_201252.Monochromatic light curves of the two meteors are shown in Fig.26. The orbital data for both these meteoroids (Table 4) indicatean origin in the main asteroid belt (TJ = 3.27 and TJ = 3.88respectively). None of the iron or Fe-rich meteors in our samplewere assigned to a known meteoroid stream. Similarly, all of theiron meteoroids detected by Borovicka et al. (2005) and Vojáceket al. (2019) were sporadic.

We also assume the possibility that Fe-rich spectra repre-sent meteoroids similar to stony iron meteorites. These types ofbodies formed either in the transition region between the coreand the mantle of differentiated asteroids (pallasites) (Norton2002), or by low-velocity collisions of large differentiated as-teroids (mesosiderites) (Wasson & Rubin 1985). The best candi-date for a stony iron meteoroid is meteor M20160402_020222.The spectral profile of this meteor can be seen in Fig. 25. Themeteor exhibited similar spectral features from all 15 frames ofthe recording. The trajectory of the meteor started low and wasshort, with a smooth light curve. These characteristics are notunlike iron meteors. The main spectral difference is the presenceof the Na I line and a possible contribution of Mg I, which ishowever likely blended with surrounding Fe I lines. The lines at558.9 nm and 616.3 nm are well fitted by Ca I. This could beexplained by the Ca-rich plagioclase grains commonly found inmesosiderites (Papike 1998). All other unmarked lines are most

Article number, page 17 of 19

Page 18: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

A&A proofs: manuscript no. AA_2019_36093_p

(a) (b)

Fig. 26: Monochromatic light curves of Fe-rich meteorsM20150508_002949 (a) and M20170301_201252 (b) showingmultiple flares. Here, the Fe I monochromatic light curve is onlyrepresented by the intensity of the 526.9 nm line. One frame cor-responds to a time period of 1/12 s.

probably caused by Fe I and FeO with a possible contributionof CaO. A detailed study of this spectral type captured in higherresolution is needed to confirm the corresponding composition.

Meteor M20160402_020222 was not captured by multiplestations of the AMOS network, so the trajectory and orbit couldnot be determined from our data. To obtain some informationabout the source of this body, the meteor was searched for in theEDMOND (European viDeo MeteOr Network Database) (Ko-rnoš et al. 2014). According to EDMOND, the meteor had a lowvelocity (vi = 14 km s-1), low beginning height (HB = 74.5 km),and low terminal height (HE = 54.4 km). This corresponds toan Apollo-type asteroidal orbit (a = 1.8 au, i = 10.4 deg) withmedium eccentricity (e = 0.46). Since numerous networks con-tribute to EDMOND, the precision of the data can vary from caseto case.

7.3. Na-enhanced and Na-rich

Na-enhanced and Na-rich meteors constitute a relatively largepart of our sample (Fig. 7). Above, we mentioned the detectedoverall enhancement of the Na/Mg intensity ratio compared tosurveys of fainter meteors. The overall results of this work showthat sodium is better preserved in larger meteoroids (Fig. 8).The study of Vojácek et al. (2019) suggests that the preserva-tion of Na is particularly efficient for meteoroids consisting oflarger grains. Considering the differences between the size popu-lations, we individually evaluated each meteor before it was clas-sified as Na-enhanced or Na-rich, depending on meteor speedand position within the ternary diagram. The contribution of Na-enhanced and Na-rich meteors is still notably higher than foundamong smaller meteoroids studied by Borovicka et al. (2005).Our sample includes 15 Na-enhanced and 4 Na-rich meteors, ofwhich 10 and 3 were observed by multiple stations, respectively.

As apparent from Fig. 8 and similarly to the results ofBorovicka et al. (2005), we found a strong connection betweenthe increase of Na/Mg intensity and meteor speed. While theeffect is taken into account for the spectral classification, the dis-tinction between normal-type and Na-enhanced (Na-rich) mete-ors is particularly difficult for very slow meteors. Various de-grees of Na enhancement are already apparent for all meteorswith entry speed vi < 27 km s-1. Meteors with vi < 15 km s-1

were identified mainly as Na-rich, Na-enhanced, or Fe-rich.The distinct spectra of Na-rich meteors raised the possibility

of atypical meteoroid composition, possibly related to inhomo-geneous structures in comet interiors. Previous surveys of fainter

meteors however consisted of only limited samples of Na-richmeteoroids, specifically lacking orbital data. Based on our re-sults, the Na enhancement detected in slow meteors is mainly, ifnot exclusively, caused by the speed preference of the low ex-citation Na I line. Essentially, the effect reflects the lower tem-perature of the radiating plasma (Borovicka et al. 2005) in com-bination with typically moderate magnitudes of these meteors.Still, two Na-enhanced meteors were detected with high me-teor speeds of ≈ 60 km s-1. The meteoroids were members ofthe Perseid and 49 Andromedid stream. The two samples con-firm that real compositional Na enhancement can be present inmeteoroids.

Based on the determined orbital and structural properties, wereport that the Na-enhanced class is a heterogeneous group ofboth cometary and asteroidal meteoroids, which are not linkedby composition. Na-rich meteors have all been found to be as-teroidal, moving on heliocentric orbits very similar to severalApollo asteroids. The asteroidal orbits correspond well withthe determined high material strengths characteristic of ordi-nary or carbonaceous chondrites (Fig. 20). We believe this spec-tral group consists of chondritic fragments of S-type or C-typeApollo asteroids. The full analysis including higher-resolutiondata collected from global AMOS stations will be presented sep-arately in Matlovic et al. (2019, in preparation).

Conclusions

We present the first wide survey of spectral and physical proper-ties of mm- to dm-sized meteoroids. While the specific materialtypes (e.g., types of chondrites) cannot be distinguished from thelow-resolution emission spectra (Borovicka et al. 2015; Drouardet al. 2018), the combination of spectral, orbital, and atmosphericmeteor data allows us to reveal the rough composition, structure,and source of meteoroids.

The obtained results reveal the compositional differences be-tween cm- to dm-sized and mm-sized meteoroids. The preser-vation of volatiles in larger meteoroids is directly observed. Wefind an overall increase of Na/Mg ratio in cm- to dm-sized mete-oroids compared to the population of mm-sized bodies. We sug-gest that this distinction is caused by the weaker effects of spaceweathering (solar radiation, solar wind and cosmic rays), whichless efficiently alter the bulk composition of larger meteoroids.

We detect a very low abundance of iron meteoroids in oursample. This is in strong contrast to mm-sized bodies, amongwhich irons dominate the meteoroids on asteroidal orbits. Ourresults suggest that most cm- to dm-sized meteoroids on aster-oidal orbits are chondritic. The large ratio of iron meteoroids inthe mm-sized population could relate to the specific formationprocess of these bodies and typical sizes of iron–nickel grains.Moreover, a new spectral group was discovered - Fe-rich me-teors, which show a higher intensity of Fe compared to mostchondritic meteoroids. We suggest that this group may containiron-enhanced H-type chondrites and bodies similar to stony ironmeteorites.

Thermal desorption during close perihelion approaches wasconfirmed as the main cause of Na depletion in medium-sizedmeteoroids, similarly to what is observed among smaller bodies.Our results demonstrate that the alteration of meteoroids and lossof volatiles during this process leads to higher material strength.A significant contribution of Na-enhanced and Na-rich mete-oroids was detected, with spectra in most cases affected by lowmeteor speed and moderate magnitudes. Some Na-enhanced andall Na-rich meteoroids were identified as asteroidal. A definitecompositional Na enhancement was detected in two cometary

Article number, page 18 of 19

Page 19: Spectral and orbital survey of medium-sized meteoroids - arXiv · 2019-08-27 · 3 we know that meter-scale meteoroids can be compositionally heterogeneous (Jenniskens et al. 2009;

Pavol Matlovic et al.: Spectral and orbital survey of medium-sized meteoroids

meteors of Perseids and 49 Andromedids, which had surpris-ingly high material strengths.

Several major meteor showers were identified within thiswork, revealing significant heterogeneity in their meteoroidstreams. We have found that Perseids originating in comet109P/Swift-Tuttle contain a wide range of materials from Na-poor to Na-enhanced meteoroids. Significant structural andspectral differences were also found among the ecliptical α-Capricornids of comet 169P/NEAT and sungrazing δ-Aquaridsof comet 96P/Machholz. The detected heterogeneities are af-fected by environmental factors and thermal history, but likelyalso reflect real inhomogeneities in comet interiors. Further-more, we find spectral similarity between κ-Cygnids and Taurids,which could imply a similar composition of the parent objects ofthe two streams.

The heterogeneity of materials originating in one parent ob-ject can also be significant for asteroidal bodies, as was sug-gested from the Príbram and Neuschwanstein meteorite pair(Spurný et al. 2003) and observed in the variety of samples foundin the Almahata Sitta (Jenniskens et al. 2009) or Benešov falls(Spurný et al. 2014). Regarding sungrazing meteoroids, the dy-namic age of these bodies is presented by the degree of Na deple-tion in their spectra. For the α-Capricornids, diverse light curvesand emission spectra reflect a structurally heterogeneous streamwith a high volatile content. We also present the first spectraldata for individual meteors originating in several minor mete-oroid streams.Acknowledgements. We are thankful to Dr. Robert Jedicke for his valuable re-view. The first author would like to thank Dr. Jirí Borovicka, Dr. Pavel Spurnýand Dr. Mária Hajduková for their comments and suggestions, which helped im-prove this work. This work was supported by the Slovak Research and Develop-ment Agency under the contract APVV-16-0148 and by the Slovak Grant Agencyfor Science grant VEGA 01/0596/18. This work was also supported by theComenius University grant UK/179/2018 and special scholarship of the Facultyof Mathematics, Physics and Informatics, Comenius University in Bratislava.

ReferencesBertaux, J. L., Lallement, R., Ferron, S., Boonne, C., & Bodichon, R. 2014,

Astronomy & Astrophysics, 564, A46Bloxam, K. & Campbell-Brown, M. 2017, Planet. Space Sci., 143, 28Borovicka, J. 1993, Astronomy & Astrophysics, 279, 627Borovicka, J. 2001, in ESA Special Publication, Vol. 495, Meteoroids 2001 Con-

ference, ed. B. Warmbein, 203–208Borovicka, J. 2010, in Proceedings of the International Meteor Conference, 26th

IMC, Bareges, France, 2007, ed. J. Rendtel & J. Vaubaillon, 42–51Borovicka, J., Koten, P., Spurný, P., Bocek, J., & Štork, R. 2005, Icarus, 174, 15Borovicka, J., Spurný, P., & Brown, P. 2015, Small Near-Earth Asteroids as a

Source of Meteorites, ed. P. Michel, F. E. DeMeo, & W. F. Bottke, 257–280Borovicka, J., Spurný, P., & Keclikova, J. 1995, Astronomy & Astrophysics Sup-

plement, 112, 173Brownlee, D., Joswiak, D., & Matrajt, G. 2012, Meteoritics and Planetary Sci-

ence, 47, 453Campbell-Brown, M. 2015, Planetary and Space Science, 118, 8Ceplecha, Z. 1968, SAO Special Report, 279Ceplecha, Z. 1987, Bulletin of the Astronomical Institutes of Czechoslovakia,

38, 222Ceplecha, Z. 1988, Bulletin of the Astronomical Institutes of Czechoslovakia,

39, 221Ceplecha, Z. & McCrosky, R. E. 1976, Journal of Geophysical Research, 81,

6257Ceplecha, Z., Spurny, P., Borovicka, J., & Keclikova, J. 1993, Astronomy &

Astrophysics, 279, 615Drouard, A., Vernazza, P., Loehle, S., et al. 2018, Astronomy & Astrophysics,

613, A54Ferus, M., Koukal, J., Lenža, L., et al. 2018, Astronomy & Astrophysics, 610,

A73Girin, O. G. 2017, Astronomy & Astrophysics, 606, A63Granvik, M., Morbidelli, A., Vokrouhlický, D., et al. 2017, A&A, 598, A52Jenniskens, P. 2006, Meteor Showers and their Parent Comets

Jenniskens, P., Gural, P., & Berdeu, A. 2014, Meteoroids 2013, 117Jenniskens, P., Nénon, Q., Albers, J., et al. 2016a, Icarus, 266, 331Jenniskens, P., Nénon, Q., Gural, P. S., et al. 2016b, Icarus, 266, 355Jenniskens, P., Shaddad, M. H., Numan, D., et al. 2009, Nature, 458, 485Jenniskens, P. & Vaubaillon, J. 2008, AJ, 136, 725Jenniskens, P. & Vaubaillon, J. 2010, Astronomical Journal, 139, 1822Jopek, T. J. & Kanuchová, Z. 2017, Planetary and Space Science, 143, 3Kasuga, T., Yamamoto, T., Kimura, H., & Watanabe, J. 2006, Astronomy & As-

trophysics, 453, L17Kohout, T., Kiuru, R., Montonen, M., et al. 2011, Icarus, 212, 697Kornoš, L., Matlovic, P., Rudawska, R., et al. 2014, Meteoroids 2013, 225Kornoš, L., Duriš, F., & Tóth, J. 2015, in International Meteor Conference Mis-

telbach, Austria, ed. J.-L. Rault & P. Roggemans, 101–104Kornoš, L., Duriš, F., & Tóth, J. 2018, in International Meteor Conference Pet-

nica, Serbia, ed. M. Gyssens & J.-L. Rault, 46–49Koten, P., Borovicka, J., Spurný, P., Betlem, H., & Evans, S. 2004, Astronomy

& Astrophysics, 428, 683Kramida, A., Yu. Ralchenko, Reader, J., & and NIST ASD Team.

2018, NIST Atomic Spectra Database (ver. 5.6.1), [Online]. Available:https://physics.nist.gov/asd [2019, March 4]. National Institute ofStandards and Technology, Gaithersburg, MD.

Kresak, L. 1979, Dynamical interrelations among comets and asteroids, ed.T. Gehrels, 289–309

Madiedo, J. M., Ortiz, J. L., Trigo-Rodríguez, J. M., et al. 2014a, Icarus, 231,356

Madiedo, J. M., Trigo-Rodríguez, J. M., Konovalova, N., et al. 2013, MonthlyNotices of the Royal Astronomical Society, 433, 571

Madiedo, J. M., Trigo-Rodríguez, J. M., Ortiz, J. L., Castro-Tirado, A. J., &Cabrera-Caño, J. 2014b, Icarus, 239, 273

Matlovic, P., Tóth, J., & Kornoš, L. 2019, Icarus, in preparationMatlovic, P., Tóth, J., Rudawska, R., & Kornoš, L. 2017, Planetary and Space

Science, 143, 104Moore, C. E. 1945, Contributions from the Princeton University Observatory, 20,

1Norton, O. R. 2002, The Cambridge Encyclopedia of Meteorites, 374Papike, J. J. 1998, Planetary materials, Vol. 36 (Mineralogical Society of Amer-

ica)Rudawska, R., Tóth, J., Kalmancok, D., Zigo, P., & Matlovic, P. 2016, Planetary

and Space Science, 123, 25Rudawska, R., Zender, J., Jenniskens, P., et al. 2014, Earth Moon and Planets,

112, 45SonotaCo. 2009, WGN, Journal of the International Meteor Organization, 37, 55Southworth, R. B. & Hawkins, G. S. 1963, Smithsonian Contributions to Astro-

physics, 7, 261Spurný, P., Haloda, J., Borovicka, J., Shrbený, L., & Halodová, P. 2014, A&A,

570, A39Spurný, P., Oberst, J., & Heinlein, D. 2003, Nature, 423, 151Tóth, J., Kornoš, L., Zigo, P., et al. 2015, Planetary and Space Science, 118, 102Tóth, J., Šilha, J., Matlovic, P., et al. 2019, in 1st NEO and Debris Detection

Conference, Darmstadt, Germany, ed. T. Flohrer, R. Jehn, & F. Schmitz, inpress

Trigo-Rodriguez, J. M., Llorca, J., Borovicka, J., & Fabregat, J. 2003, Meteorit-ics and Planetary Science, 38, 1283

Trigo-Rodriguez, J. M., Madiedo, J. M., Williams, I. P., & Castro-Tirado, A. J.2009, MNRAS, 392, 367

Tubiana, C., Snodgrass, C., Michelsen, R., et al. 2015, Astronomy & Astro-physics, 584, A97

Capek, D. & Borovicka, J. 2017, Planetary and Space Science, 143, 159Vida, D., Brown, P. G., & Campbell-Brown, M. 2018, MNRAS, 479, 4307Vojácek, V., Borovicka, J., Koten, P., Spurný, P., & Štork, R. 2015, Astronomy

& Astrophysics, 580, A67Vojácek, V., Borovicka, J., Koten, P., Spurný, P., & Štork, R. 2019, Astronomy

& Astrophysics, 621, A68Wasson, J. T. & Rubin, A. E. 1985, Nature, 318, 168Wojdyr, M. 2010, Journal of Applied Crystallography, 43, 1126Zigo, P., Toth, J., & Kalmancok, D. 2013, in Proceedings of the International

Meteor Conference, 31st IMC, La Palma, Canary Islands, Spain, 2012, ed.M. Gyssens & P. Roggemans, 18–20

Article number, page 19 of 19